We compute the heat kernel coefficients that are needed for the regularization and renormalization of massive gravity. Starting from the Stueckelberg action for massive gravity, we determine the propagators of the different fields (massive tensor, vector and scalar) in a general linear covariant gauge depending on four free gauge parameters. We then compute the non-minimal heat kernel coefficients for all the components of the scalar, vector and tensor sector, and employ these coefficients to regularize the propagators of all the different fields of massive gravity. We also study the massless limit and discuss the appearance of the van Dam–Veltman–Zakharov discontinuity. In the course of the computation, we derive new identities relating the heat kernel coefficients of different field sectors, both massive and massless.
I. INTRODUCTION
In both mathematics and physics the heat kernel technique is a well-established method for the computation of traces of differential operators.1–6 Its area of application ranges from fluctuations of quantum fields on curved spacetimes, the determination of ultraviolet divergences, effective actions and quantum anomalies to quantum gravity and further. Concretely, it allows to define the regularized propagator and using this the one-loop effective action, and to compute counterterms and anomalies in a way that is naturally extended to field theories on curved space. Furthermore, it constitutes an essential ingredient in solving the gravitational functional renormalization group equation.7,8
The interest in the heat kernel can be traced back to the beginning of the theory of quantum fields in curved spacetimes. In the 1950s, Morette and DeWitt established the path integral approach to propagators in curved spacetime, and in the 1960s Schwinger and DeWitt developed the proper time formalism and the associated heat kernel expansion. The main idea is to express a given Green’s function as an integral of the heat kernel (which satisfies the heat equation with time τ, the proper time) over the proper time. The heat kernel depends on the background geometry, especially the background metric, and possibly other background fields such as gauge fields. Except for some special manifolds, such as maximally symmetric spacetimes, it is not possible to compute it exactly in general. However, it is possible to compute it approximately, and up to now two main expansion schemes have been developed: the small-time or local expansion (also known as the Seeley–DeWitt technique), and a non-local expansion for small curvature. The first scheme is employed in quantum field theory to compute ultraviolet divergences and anomalies, since these are directly determined by the lowest-order heat kernel coefficients in the small-time expansion. The non-local expansion is used instead to compute the finite part of the effective action in a covariant manner. In the framework of quantum gravity, non-local heat kernels are crucial in renormalization group studies, such as tests of the Asymptotic Safety conjecture, and heat kernels on maximally symmetric spaces are used to find β functions for gravitational couplings.9–12 A recent development in this framework is the consideration of Lorentzian renormalization group (RG) flows,13–18 which is important since a generic metric does not admit a Wick rotation. In general, including couplings with increasing scaling dimension such as higher curvature terms, an increasing number of heat kernel coefficients is needed, whose computation becomes increasingly difficult.
For gauge theories of Yang–Mills type it is straightforward to determine classical observables using the Becchi–Rouet–Stora–Tyutin (BRST) formalism.19–21 Since these generally are not linear in the elementary fields, in the quantum theory they become composite operators which need additional renormalization, beyond the usual renormalization of couplings.22 This can also be done using the heat kernel technique, point-splitting the classical expression and subtracting as many terms of the small-time expansion of the heat kernel as are needed to obtain a finite coincidence limit (depending on the scaling dimension of the operator). When using an effective action, one has to introduce an additional term coupling the composite operator to an external source, renormalize the extended effective action, and then take functional derivatives with respect to the source to obtain correlation functions including insertions of this operator. In the exact renormalization group framework, this has been recently used to study flows of volume and length operators in gravity,23–25 see also Refs. 26 and 27. For example, when studying the correlator of metric fluctuations in the vielbein formalism, the metric itself is a composite operator which needs to be regularized and renormalized in a suitable way, such as the point-splitting method28–30 or the heat kernel renormalization method explained above.
However, when defining observables in gravity another complication arises. Namely, in contrast to Yang–Mills gauge theories where the gauge symmetry is an internal symmetry and acts locally, i.e., it transforms fields at the same point, the diffeomorphism symmetry of gravity moves points around. It follows immediately that a local field cannot be gauge-invariant, and that observables must be necessarily nonlocal. Various approaches to construct such nonlocal observables have been considered, for example defining correlation functions involving the geodesic distance between points31–42 or relational observables. Nevertheless, once one has suitable observables at hand, also for gravity the usual framework of regularization and renormalization applies, and a natural question that arises is the gauge dependence of those regularized and renormalized quantities in the quantum theory. While the gauge independence of the observables is clear in the classical theory, in the quantum theory anomalies may arise which then might render the final result gauge-dependent. To check the gauge independence of the renormalized observables, it is useful to work in a general family of gauges that depend on a certain number of gauge parameters, and then verify that the end result is independent of these parameters. For this, it is necessary to know the Green’s functions and the heat kernel coefficients in such a general family of gauges. In Ref. 43, the authors constructed Green’s functions in a globally hyperbolic spacetime in general linear covariant gauges, both for vector gauge bosons and linearized Einstein gravity in the presence of a cosmological constant. In this work, we will generalize their result to the massive gravity case, relating the propagators of the massive gravity field sectors in a general linearized gauge to their expressions in the corresponding Feynman gauge.
The choice of the massive gravity theory is motivated from both a phenomelogical and cosmological point of view, and for theoretical interest. Modifications of Einstein gravity, of which massive gravity theories are a subset, give rise to an effective cosmological fluid which can explain dark matter and/or dark energy.44–46 On the other hand, while there are stringent upper bounds on the possible mass of the graviton,47,48 a tiny mass is not ruled out completely, even in light of the van Dam–Veltman–Zakharov (vDVZ) discontinuity.49,50 On the theoretical side, to the best of our knowledge the propagators for a massive spin-2 field in a general gauge have not been determined, and consequently also the corresponding heat kernel expansion is unknown. To avoid breaking diffeomorphism invariance, which a simple Fierz–Pauli mass term51–53 would do, we use the Stueckelberg formalism54,55 and add additional spin-1 and spin-0 fields that transform in such a way to keep diffeomorphism invariance of the complete theory. That is, we consider in this work only the free massive gravity theory, which determines the propagators. Moreover, we work with an on-shell background satisfying the background Einstein equations, which means that the propagators that we derive could, for example, later on be used to determine an on-shell effective action for gravity.56 While this is not strictly necessary, since one could also work with the off-shell (Vilkovisky–deWitt) invariant effective action,57 the resulting expressions are already complicated enough for an on-shell background, and would become unmanageable off-shell.
This article is structured as follows: In Sec. II we give some more background on massive gravity, determine its linearized action including the Stueckelberg fields that are necessary to preserve diffeomorphism invariance, and using the BRST formalism add the gauge-fixing and ghost terms that determine a quite general linear covariant gauge for the fields. Section III is devoted to the determination of the propagators in this gauge, which comprises three free gauge parameters, as well as their massless limit, and the verification of the Ward identities in the free theory. In Sec. IV, we give an overview of the small-time expansion of the heat kernel, compute the leading three heat kernel coefficients, and determine relations between the heat kernel coefficients of different spins. Section V comprises our main results, the heat kernel expansion of the propagators for all fields in massive gravity, including the massless limit. Finally, in Sec. VI we discuss our result and present an outlook for future work.
Conventions: Our conventions are a mostly plus metric, the Riemann tensor defined by ∇μ∇νvρ − ∇ν∇μvρ = Rμνρσvσ, and the Ricci tensor . We work in n dimensions, use geometrical units ℏ = c = 1, and set with Newton’s constant GN and the Planck length ℓPl.
II. MASSIVE GRAVITY
Even though General Relativity (GR) has been very well tested and successfully describes gravity both a small and large scales,58 there are also some tensions. Of those, let us mention the observed galaxy rotation curves59,60 and primodial inflation,61–64 which cannot be explained using GR alone. Various models have been proposed to explain these effects and alleviate the tensions, of which dark matter–matter that only interacts gravitationally, or only very weakly with Standard Model particles — can explain the galaxy rotation curves and dark energy — an effective fluid with negative pressure — can explain inflation.64 While dark matter can be easily added as a new fundamental particle (or several) such as axions, and the current expansion can be explained by a cosmological constant, primordial dark energy is usually modelled by a single fundamental scalar field (the inflaton) with a suitable potential, see Refs. 65 and 67 for recent reviews. Other fundamental explanations of dark matter and dark energy come in the form of modified gravity models,68 of which massive gravity in its various incarnations45,69–72 is a popular one.
In massive gravity, the mediator of gravitational interactions, the spin-2 field usually called graviton, is given a small mass. From a particle physics perspective, this is a natural modification since GR can be thought of as the unique theory of a massless spin-2 particle;73–75 it also seems natural since we know that the carrier particles of the electroweak forces acquire a mass through the Higgs mechanism. Since the force mediated by massive particles falls off exponentially at large distances, a graviton with a small mass of the order of the Hubble rate m ∼ H could also mimic the effects of the current accelerated expansion of our universe without invoking a cosmological constant or other forms of dark energy. Another motivation for infrared modifications of gravity comes from brane world models,76,77 where extra dimensions can be of large or infinite size and the effective four-dimensional graviton propagator on our brane can behave like a massive one. We note that giving a mass to a free spin-2 field is easy, and was already done in 1939 by Fierz and Pauli;51–53 the construction of interacting theories of massive gravity is much harder. One possibility for a non-linear generalization of the Fierz–Pauli mass term was explored in Ref. 78, where an additional fully dynamical “reference” metric was introduced, rendering the model invariant under general coordinate transformations. Such theories are called bi-metric theories of gravity, and have been of particular interest due to the presence of accelerating cosmological solutions. Observational data from both the early evolution of the universe and solar system tests of gravity can be used to constrain the parameters of these theories.79–82 Another possibility is topologically massive gravity83,84 and its generalizations,85 which was the first interacting model of massive gravity in three dimensions. Its β functions can be computed in the asymptotic safety approach, and it turns out that as Einstein gravity in four dimensions,9,10 topologically massive gravity is asymptotically safe.86 Since Lorentz-invariant models of massive gravity often suffer from the presence of ghosts (fields with a wrong sign of the kinetic term), a possibility which recently also attracted attention is to allow for a violation of Lorentz invariance.87–91 From the cosmological point of view, this is particularly interesting because one can construct a consistent model of gravity where the tensor graviton mode is massive, while the linearized equations for scalar and vector metric perturbations are not modified. The Friedmann equation then acquires an effective dark-energy component, leading to an accelerated expansion, while gravity in the solar system is not modified.
However, in this work we are interested in the linearized Fierz–Pauli theory that preserves Lorentz invariance, and which can be obtained as the first approximation of a general massive gravity theory around a given fixed background. It turns out that even in the massless limit, the extra polarizations of a massive spin-2 field do not completely decouple in general, unlike what happens for an Abelian gauge theory. Instead, a scalar mode remains coupled to the tensor modes, and the resulting theory differs from the completely massless one, a effect that is known as the van Dam–Veltman–Zakharov (vDVZ) discontinuity.49,50,92,94 Nevertheless, whether decoupling occurs depends strongly on the concrete theory, which in our case means the background around which we study the theory. Let us mention here that the discontinuity is absent when one considers GR with a non-vanishing cosmological constant Λ. While for an anti-de Sitter background with Λ < 0 the resulting theory is ghost-free for all masses,94,95 in n-dimensional de Sitter spacetime with Λ = (n − 1)(n − 2)/2H2 > 0 the theory has a scalar ghost if 0 < m2 < (n − 2)H2; the upper bound (n − 2)H2 is known as the Higuchi bound.96 For a modern account of these issues we refer the reader to Refs. 97–99 and references therein.
Even though Fierz–Pauli theory preserves Lorentz invariance, the mass term breaks linearized diffeomorphisms, which are a symmetry of the massless theory that descends from the full Einstein–Hilbert action for gravity. To remedy this, we employ the Stueckelberg trick54,55 and add auxiliary fields that transform in such a way as to keep the full action invariant under linearized diffeomorphisms. The original massive Fierz–Pauli theory can then be recovered in a special gauge, the so-called unitary gauge. While both perturbative and non-perturbative computations in gravity have used heat kernel expansions (see for example Ref. 100 for a recent study), and also massive vector fields have been studied in great detail,101–103 to our knowledge the present work represents the first computation of the heat kernel coefficients for a massive theory of gravity in a general gauge.
A. Action
B. BRST, gauge fixing and ghosts
To quantize the Stueckelberg gauge theory of (free) massive gravity with action (11), we use the standard BRST formalism,19–21 where we introduce ghost, antighost and auxiliary fields for every symmetry. The gauge transformations (10a) are replaced by the action of the fermionic BRST operator whose action on fields is obtained by replacing the gauge parameter by the corresponding ghost field, and the action on the ghost, antighost and auxiliary fields must be determined such that the action is nilpotent: .
III. PROPAGATORS
To determine the propagators of the different fields in a general gauge, we have to relate them to the propagators in Feynman-type gauges, where the corresponding equation of motion is normally hyperbolic. In normally hyperbolic equations of motion, second-order derivatives only appear via the d’Alembertian ∇2, and their (retarded and advanced) propagators or Green’s functions can be constructed in an arbitrary globally hyperbolic spacetime using successive approximations.105–107 As we will see below, in a general gauge where second-order derivatives appear also in a different form the propagators can be obtained as a linear combination of the Feynman-gauge propagator and derivatives of propagators for fields of lower spin. For this, we will first derive various identities for the Feynman-gauge propagators, then treat the ghost sector where a spin-1 propagator in a general gauge is needed, and afterwards consider the fields sector where in addition also the spin-2 propagator in a general gauge appears. To derive these identities, we will make repeated use of the fact that the solution of a normally hyperbolic equation with retarded or advanced boundary conditions is unique, in particular that such an equation with vanishing source only has a vanishing solution. Lastly, we verify that our propagators satisfy the relevant Ward(–Takahashi–Slavnov–Taylor) identities.
A. Feynman-type gauges
B. The ghost sector
C. The field sector
D. Auxiliary fields
E. Ward identities
Ward(–Takahashi–Slavnov–Taylor) identities116–119 are relations between correlation functions that stem from the gauge invariance of the underlying classical theory. In the BRST formalism, they arise from the fact that any BRST-exact term has vanishing expectation value in a physical state, i.e., . In our case, they relate the different propagators of the fields, ghosts and auxiliary fields, and we will display explicitly a subset of all possible identities.
IV. HEAT KERNEL
The heat kernel technique was introduced in quantum field theory as a way to treat functional traces and determinants of local differential operators of Laplace type, i.e., second-order elliptic differential operators whose principal symbol is the metric (also called minimal operators). Later on, it was generalized to also include non-minimal operators, where a more general principal symbol is admissible, and which arise for example in gauge theories in a gauge different from Feynman gauge, as well as hyperbolic operators. In the following we will give a short overview of the technique, and refer to the reviews and books1,2,5,120–122 for more in-depth results.
For non-minimal operators, there is no direct heat kernel expansion. Unfortunately, many of the operators that are of physical interest are not of Laplace-type. A well-known example is Yang-Mills theory in a covariant gauge with parameter ξ different from Feynman gauge ξ = 1, where the bundle is composed from the tangent bundle and an internal Lie algebra bundle (depending on the gauge group). While for a semi-simple Lie algebra the internal part of PAB (the Cartan–Killing form) is proportional to the identity δab, the covariant spacetime derivatives are uncontracted, and is a non-minimal differential operator. One possibility to treat such operators is the use of covariant projectors which lead to Laplace-type differential operators acting on subspaces of the original field space, i.e., on a different vector bundle. Typical examples for such projections are the transverse decomposition of a vector field, or the transverse traceless decomposition of a symmetric tensor field, for which one can in particular use the second approximation scheme explained above.129–132 Alternatively, one can write the heat kernel for the non-minimal operator as the sum of the one for the minimal operator and covariant derivatives of the heat kernel of lower spin fields as in Ref. 2, Sec. 2. This is akin to the determination of the corresponding Green’s functions in terms of the Feynman-type gauge ones, and is based on the Ward identities that hold in the free theory analogously to the trace and divergence identities that we computed in Sec. III. In fact, we will derive below trace and divergence identities for the (minimal, Feynman-type gauge) heat kernels (and their coefficients), using which the corresponding non-minimal heat kernel could be computed. Unfortunately, the relation between the two approaches to non-minimal operators and their heat kernel expansions is not straightforward, mainly because the known expansions are in general only asymptotic. One can obtain a convergent expansion by multiplying the coefficients in the local expansion (98) by cutoff functions whose support decreases quickly enough with k (see Ref. 133, Sec. 5.2). In this way, the same coincidence limits (as x′ → x) as for the unmodified expansion hold, but in any finite neighborhood the sum only has a finite number of terms; only for analytic spacetimes one can hope to have a convergent local expansion without cutoffs. Fortunately, for most purposes it is enough to know the coincidence limits of a finite number of coefficients and their derivatives, which are always well-defined.
The main application of the heat kernel in quantum field theory is in the renormalisation of the effective action or of composite operators, for which only the coincidence limit of a finite number of coefficients and their derivatives are needed. In particular, for the renormalisation of the stress tensor in n dimensions one needs the coincidence limit of the coefficients Ak with k = 0, …, n. To check that the renormalized stress tensor is covariantly conserved, one needs in addition the coincidence limit of their derivatives with k + ℓ ≤ n + 1. To determine these coincidence limits, one can either use the explicit formula (109) in normal coordinates, or one can take covariant derivatives of the transport Eq. (102) and their coincidence limit. In turn, these can be obtained recursively using the coincidence limits of derivatives of the Synge world function σ and the van Vleck–Morette determinant Δ.
Several methods have been developed to find explicit expressions for the heat kernel coefficients.137 DeWitt28 determined the first two coefficients with a covariant recursive method. Sakai134 relied on the Riemann normal coordinates to find the third coefficient in the scalar case (i.e., for a single scalar field on a curved space). For the general case, this coefficient was found by Gilkey138 using a noncovariant pseudo-differential-operator technique (see also Ref. 139). The integrated and traced fourth and fifth coefficients for an arbitrary field theory in flat space were found in Ref. 140 through the evaluation of a noncovariant Feynman graph. Avramidi141,142 presented a new covariant non-recursive procedure and found the fourth coefficient for the general case; the coefficient a5 in flat space was computed in Refs. 140 and 143. Various resummation methods have also been developed, see Ref. 144 and references therein. Finally, the connection between the Riemannian heat kernel and the Lorentzian analogue, which we treat in the following, was elucidated for stationary spacetimes by Strohmaier and Zelditch.145
A. Small time regularisation
For x ≠ x′, the factor in the heat kernel expansion (115), or in the expansion (98), ensures that the proper time integrals (100) and (112) converge at τ = 0. However, in the coincidence limit x′ = x which is needed for applications, this factor becomes 1 and these integrals are divergent for small τ. This is of course the well-known divergence of the propagators in the coincidence limit.148 In particular, it is clear from the expansions (98) and (115) that the coefficients A0, …, An/2−1 have divergent prefactors if one computes the coincidence limit of the propagator, and more if one is interested in the coincidence limit of derivatives acting on it.
To regulate this divergence and extract the finite part of the result, analytic regularisation schemes are well suited for the heat kernel expansion. A well-known such scheme is dimensional regularisation, where one continues the dimension n of spacetime into the complex plane, computes results in a region where all integrals are convergent, and then analytically continues the result back to the physical dimension. In particular, for small enough (depending on the number of covariant derivatives acting on the heat kernel), the proper time integrals are convergent for small τ even in the coincidence limit. The original divergence of the propagator then manifests itself in poles as n approaches the physical dimension. While dimensional regularisation is versatile and usually the simplest one, it also has problems, namely when the objects under study cannot naturally be defined in arbitrary dimension n. In particular, this concerns chiral, conformal and supersymmetric theories, where the symmetries are dimension-dependent. These symmetries can then be broken in the quantum theory and one obtains chiral, conformal and supersymmetry anomalies.149,150 Of course anomalies are physical and cannot depend on the choice of a regularisation scheme, but only on renormalisation conditions (as can be seen in the dichotomy between covariant and consistent anomalies151). Therefore, for dimension-dependent symmetries where the advantages of dimensional regularisation are absent, it can be easier to employ a different regularisation scheme. For the heat kernel, such a scheme is obtained by inserting a factor τδ in the integrals (98) and (115) and performing the analytic continuation in δ.
B. Generic bundle
In the following sections, we specialize these results to scalar, vector and tensor coefficients.
C. Scalar coefficients
D. Vector coefficients
The coefficients again agree with results in the literature (see for example Ref. 128).
E. Tensor coefficients
F. Relations between coefficients
From the trace and divergence identities for the vector and tensor propagators in Feynman-like gauges, we obtain relations between their corresponding heat kernel coefficients.
V. HEAT KERNEL EXPANSION OF PROPAGATORS
Having obtained all the ingredients, the heat kernel expansion can be used to derive the regularized propagators of a massive theory of gravity. In Sec. III we obtained the propagators in the various fields sectors, which we now express using the heat kernel.