This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

aainstitutetext: Institute for Mathematics, Astrophysics and Particle Physics (IMAPP), Radboud University, Heyendaalseweg 135, 6525 AJ Nijmegen, The Netherlandsbbinstitutetext: Theory Department, Lebedev Physics Institute, Leninsky Prospect 53, Moscow 119991, Russia

On the origin of almost-Gaussian scaling
in asymptotically safe quantum gravity

Maximilian Becker,111Corresponding author. b    Alexander Kurov a    and Frank Saueressig [email protected] [email protected] [email protected]
Abstract

The gravitational asymptotic safety program envisions the high-energy completion of gravity through an interacting renormalization group fixed point, the Reuter fixed point. The predictive power of the construction is encoded in the spectrum of the stability matrix which is obtained from linearizing the renormalization group flow around this fixed point. A key result of the asymptotic safety program is that parts of this spectrum exhibits an almost-Gaussian scaling behavior, entailing that operators which are classically highly UV-irrelevant do not induce new free parameters. In this article, we track down the origin of this property by contrasting the structure of the stability matrix computed from the Wetterich equation and the Composite Operator equation within the realm of f(R)f(R)-truncations. We show that the almost-Gaussian scaling is not linked to the classical part of the beta functions. It is a quantum-induced almost-Gaussian scaling originating from the quantum corrections in the flow equation. It relies on a subtle interplay among the analytic structure of the theory’s two-point function and the way the Wetterich equation integrates out fluctuation modes. As a byproduct we determine the parts of the eigenmode spectrum which is robust with respect to changing the regularization procedure.

1 Introduction

A key challenge in constructing a quantum field theory including gravitational interactions is the predictive power. This issue already appears in the perturbative quantization of the Einstein-Hilbert action where the negative mass dimension of Newton’s coupling leads to new divergences at each order in the loop expansion 'tHooft:1974bx ; Goroff:1985sz ; Goroff:1985th ; vandeVen:1991gw . Removing these divergences requires the introduction of an infinite tower of counterterms. Following the observation that “any quantum field theory is renormalizable once all possible counterterms are taken into account”, this procedure may result in a perfectly viable quantum theory. Since each counterterm introduces a new free parameter, such a construction lacks predictive power though. This underlies the assessment that general relativity is an effective field theory Donoghue:2017pgk whose high-energy completion requires new physics.

The gravitational asymptotic safety program, recently reviewed in the book chapters Saueressig:2023irs ; Pawlowski:2023gym ; Eichhorn:2022gku ; Knorr:2022dsx ; Morris:2022btf ; Martini:2022sll ; Wetterich:2022ncl ; Platania:2023srt and text books Percacci:2017fkn ; Reuter:2019byg , offers an elegant way to equip this effective field theory with predictive power. The core of the program is an interacting fixed point of the theory’s Wilsonian renormalization group (RG) flow, the Reuter fixed point Reuter:1996cp . The asymptotic safety hypothesis then states that this fixed point controls the theory as the coarse-graining scale kk is sent to infinity. The fact that this hypothesis supplies predictive power can then be seen as follows. By definition, the dependence of the dimensionless couplings {un}\{u^{n}\} on the coarse-graining scale is encoded in the beta functions

kkun(k)=βn(u).k\partial_{k}u^{n}(k)=\beta^{n}(u)\,. (1)

At a fixed point {un}\{u^{n}_{*}\} one has βn(u)=0\beta^{n}(u_{*})=0, for all nn. The RG flow in the vicinity of the fixed point can then be studied by linearizing the beta functions

kkun(k)=m(B)n(umum)m+O((ujuj)2),(B)n=mBn(u)m=βnum|u=u.k\partial_{k}u^{n}(k)=\sum_{m}(B_{*})^{n}{}_{m}(u^{m}-u^{m}_{*})+O((u^{j}-u^{j}_{*})^{2})\,,\quad(B_{*})^{n}{}_{m}=B^{n}{}{}_{m}(u_{*})=\left.\frac{\partial\beta^{n}}{\partial u^{m}}\right|_{u=u_{*}}\,. (2)

Here B(u)B(u_{*}) is the stability matrix of the fixed point. The solutions of the linearized system are readily obtained in terms of the eigenvalues λI\lambda_{I} and corresponding (right) eigenvectors VIV_{I} of B(u)B(u_{*}) and read

un(k)=un+JCJVJn(kk0)λJ.u^{n}(k)=u^{n}_{*}+\sum_{J}\,C_{J}\,V_{J}^{n}\left(\frac{k}{k_{0}}\right)^{\lambda_{J}}\,. (3)

Here, the CJC_{J} are free coefficients specifying a solution and k0k_{0} is an arbitrary reference scale. For eigendirections with Re(λJ)<0\lambda_{J})<0, the RG flow is dragged into the fixed point as kk\rightarrow\infty. Conversely, eigendirections with Re(λJ)>0\lambda_{J})>0 repel the flow in this limit. In order to reach the fixed point, the integration constants along the UV-repulsive directions must be set to zero while the ones associated with the UV-attractive directions constitute free parameters. Hence the number of free parameters is determined by the spectrum of the stability matrix and equals the number of eigenvalues λI\lambda_{I} with negative real part. More precisely, the predictive power of asymptotic safety hinges on spec(B)\mathrm{spec}(B_{*}), being bounded from below and containing a finite number of eigenvalues with ReλI<0\mathrm{Re}\,\lambda_{I}<0. Conventionally, spec(B)\mathrm{spec}(B_{*}) is given in terms of the stability coefficients, also called the critical exponents, θI\theta_{I}, defined as θIλI\theta_{I}\equiv-\lambda_{I}. Notably, spec(BB_{*}) is invariant under a redefinition of the couplings unu^{n} Reuter:2019byg . For an asymptotically free theory, it agrees with (minus) the classical mass dimension of the operator. Since the bare action contains a finite number of relevant and marginal couplings only, predictive power is guaranteed. For the asymptotic safety construction, however, the presence of quantum fluctuations makes this feature highly non-trivial.

Determining the spectrum of BB_{*} for the Reuter fixed point and its generalization to gravity-matter systems has been the subject of a vast body of research, see Lauscher:2002sq ; Benedetti:2009rx ; Benedetti:2010nr ; Benedetti:2013jk ; Dietz:2016gzg ; Gies:2016con ; Denz:2016qks ; Falls:2017lst ; Ohta:2018sze ; Sen:2021ffc ; Knorr:2021slg ; Baldazzi:2023pep for selected references. The primary approach builds on solving the Wetterich equation Wetterich:1992yh ; Morris1994 ; Reuter:1993kw ; Reuter:1996cp , a formally exact functional renormalization group equation (FRGE) encoding the Wilsonian RG flow of the effective average action. The most comprehensive studies along these lines approximate Γk\Gamma_{k} by a polynomial or a generic function in the Ricci scalar RR and determine the beta functions (1) by projecting the RG flow onto a maximally symmetric spherical background Codello:2007bd ; Machado:2007ea ; Codello:2008vh ; Benedetti:2012dx ; Dietz:2012ic ; Dietz:2013sba ; Bridle:2013sra ; Demmel:2014sga ; Demmel:2014hla ; Demmel:2015oqa ; Falls:2013bv ; Falls:2014tra ; Falls:2016wsa ; Alkofer:2018fxj ; Alkofer:2018baq ; Ohta:2015efa ; Ohta:2015fcu ; Falls:2016msz ; Mitchell:2021qjr ; Eichhorn:2015bna ; Gonzalez-Martin:2017gza ; DeBrito:2018hur ; Kluth:2020bdv ; Christiansen:2017bsy . Building on the functional f(R)f(R)-truncations Machado:2007ea ; Codello:2008vh ; Benedetti:2012dx ; Dietz:2012ic ; Dietz:2013sba , Refs. Falls:2014tra ; Falls:2018ylp considered polynomials up to order Nprop=70N_{\rm prop}=70 and observed that the eigenvalue spectrum of classically highly irrelevant operators essentially agrees with the classical scaling. This feature has been dubbed “almost-Gaussian scaling” and has been supported using various approximations of Γk\Gamma_{k} also beyond the polynomial f(R)f(R)-computations Codello:2008vh ; Gies:2016con ; Falls:2017lst ; Ohta:2018sze ; Kluth:2022vnq . Taken together, these studies suggest that the number of free parameters associated with the Reuter fixed point could be as low as 33 or 44.

Complementary to solving the Wetterich equation, the stability matrix of a fixed point can also be approximated with a composite-operator equation (COE) Pawlowski:2005xe ; Pagani:2016dof ; Pagani:2016pad . This strategy has been applied to spacetime volumes and also to operators polynomial in the Ricci scalar Houthoff:2020zqy ; Kurov:2020csd ; Martini:2023qkp .222The initial idea of the COE is that it gives access to the anomalous scaling dimension of composite operators which do not naturally appear in the effective average action. This property has been exploited to study the anomalous scaling dimension of the geodesic distance Becker:2018quq ; Becker:2019tlf ; Becker:2019fhi . The f(R)f(R)-type computations Houthoff:2020zqy ; Kurov:2020csd did not see any evidence for the almost Gaussian scaling behavior. In this case the addition of higher-order operators opened up new UV-relevant directions which is in variance with the results found by solving the Wetterich equation.

The goal of our work is to track down the origin of this mismatch. In this way, we are able to pinpoint the structural property of the Wetterich equation which underlies the almost Gaussian scaling of the operator spectrum at the Reuter fixed point. The direct comparison shows that this specific property is absent in the COE. This insight allows to tie the mismatch in the two approaches to the pole structure of the trace part of the graviton propagator on a constantly curved background. Moreover, it reveals that the “almost-Gaussian scaling” exhibited by the stability matrix is a genuine quantum phenomenon in the sense that it is not linked to the classical part of the beta functions. This analysis is an important step towards understanding one of the key properties of the gravitational asymptotic safety program, its predictive power. From a wider perspective, our results are also important when comparing the scaling dimensions computed from the composite operator equation to Monte Carlo simulations of the gravitational path integral within Euclidean Dynamical Triangulations (EDT) Laiho:2017htj and Causal Dynamical Triangulations (CDT) Ambjorn:2004qm ; Loll:2019rdj ; Brunekreef:2023ljt .

The remainder of this work is organized as follows. In Sec. 2, we derive a master equation giving the structure of the stability matrix at a generic fixed point. In Sec. 3, we compute the building blocks making up the master equation as they appear from solving the Wetterich equation and the composite operator equation for f(R)f(R)-type operators on spherically symmetric backgrounds. In Sect. 4, we link these findings to the analytic structure of the two-point function and we end with a discussion and our conclusions in Sect. 5. In order to make our work self-contained some technical details on f(R)f(R)-type truncations are given in Appendix A and B.

2 Building the stability matrix at a RG fixed point

2.1 General structure of the beta functions

The primary tool for investigating RG fixed points in quantum gravity is the effective average action Γk\Gamma_{k}, a scale-dependent generalization of the ordinary effective action Γ\Gamma. By construction, Γk\Gamma_{k} provides a description of the effective dynamics which incorporates the effects of quantum fluctuations with momenta p2k2p^{2}\gtrsim k^{2} in its effective vertices. In the context of gravity, Γk\Gamma_{k} depends on a background metric g¯μν\bar{g}_{\mu\nu} (which may be left unspecified), metric fluctuations hμνh_{\mu\nu} in this background, and ghost fields C¯μ,Cν\bar{C}_{\mu},C^{\nu}, arising from gauge-fixing the theory via the Faddeev-Popov construction. Hereby, the background metric g¯μν\bar{g}_{\mu\nu} and the fluctuations hμνh_{\mu\nu} combine to the full spacetime metric gμνg_{\mu\nu}. The presence of a fixed, non-fluctuating background is essential to be able to define the coarse-graining scale kk. The interaction monomials in Γk\Gamma_{k} are taken to be invariant under (background) diffeomorphisms. Conceptually, the gravitational effective average action may be organized according to

Γk[h,C¯,C;g¯]=Γ¯k[g]+Γ^k[h;g¯]+Γkghost[h,C¯,C;g¯],\begin{split}\Gamma_{k}[h,\bar{C},C;\bar{g}]=&\,\bar{\Gamma}_{k}[g]+\hat{\Gamma}_{k}[h;\bar{g}]+\Gamma^{\rm ghost}_{k}[h,\bar{C},C;\bar{g}]\,,\end{split} (4)

where Γ¯k[g]\bar{\Gamma}_{k}[g] is the “diagonal part” of Γk\Gamma_{k} which depends on g¯μν\bar{g}_{\mu\nu} and hμνh_{\mu\nu} in the combination of gμνg_{\mu\nu} only, Γ^k[h;g¯]\hat{\Gamma}_{k}[h;\bar{g}] is the off-diagonal part that also incorporates the gauge-fixing term, and Γkghost[h,C¯,C;g¯]\Gamma^{\rm ghost}_{k}[h,\bar{C},C;\bar{g}] is the ghost-action. In addition, we may supplement the construction by (potentially non-local) “composite operators” {𝒪~n(k)}\{{\tilde{\mathcal{O}}}_{n}(k)\} coupled to their own sources εn\varepsilon_{n},

Γk[h,C¯,C;g¯;ε]=Γk[h,C¯,C;g¯]+nεn𝒪~n(k).\begin{split}\Gamma_{k}[h,\bar{C},C;\bar{g};\varepsilon]=&\,\Gamma_{k}[h,\bar{C},C;\bar{g}]+\sum_{n}\varepsilon_{n}\,{\tilde{\mathcal{O}}}_{n}(k)\,.\end{split} (5)

This extension coincides with the standard definition (4) for εn=0\varepsilon_{n}=0.

The dependence of Γk\Gamma_{k} on the coarse-graining scale kk is governed by the Wetterich equation Wetterich:1992yh ; Morris1994 ; Reuter:1993kw ; Reuter:1996cp ,

ddtΓk=12STr[(Γk(2)+k)1ddtk].\frac{\mathrm{d}}{\mathrm{d}t}\Gamma_{k}=\frac{1}{2}\text{STr}\left[\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\,\frac{\mathrm{d}}{\mathrm{d}t}\mathcal{R}_{k}\right]\,. (6)

Eq. (6) is a functional renormalization group equation (FRGE), in which STr denotes a sum over fluctuation modes (including a sum over all fields that Γk\Gamma_{k} is a functional of) while Γk(2)\Gamma_{k}^{(2)} denotes the second functional derivative of Γk\Gamma_{k} with respect to these fluctuation fields. Moreover, k{\mathcal{R}}_{k} is a regulator that equips fluctuations of low momenta with a kk-dependent mass term, while it vanishes for fluctuations with p2k2p^{2}\gg k^{2}. Further, we introduced the RG time tln(k/k0)t\equiv\ln(k/k_{0}) for notational convenience.

Retaining the composite operators and expanding (6) to first order in εi\varepsilon_{i} gives the composite operator equation Pagani:2016pad ; Pagani:2016dof ; Becker:2018quq ; Becker:2019tlf ; Becker:2019fhi ; Houthoff:2020zqy ; Kurov:2020csd ,

ddt𝒪~n(k)=12STr[(Γk(2)+k)1𝒪~n(2)(k)(Γk(2)+k)1ddtk].\frac{\mathrm{d}}{\mathrm{d}t}\tilde{{\mathcal{O}}}_{n}(k)=-\frac{1}{2}\text{STr}\left[\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\,\tilde{{\mathcal{O}}}_{n}^{(2)}(k)\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\,\frac{\mathrm{d}}{\mathrm{d}t}\mathcal{R}_{k}\right]\,. (7)

The main virtue of this equation is that it allows to study the scaling dimension of geometric quantities which typically do not appear in Γk\Gamma_{k}. A prototypical example of the latter can be the geodesic distance between points studied in Becker:2018quq ; Becker:2019tlf . The use of (7) requires information about the propagators (Γk(2)+k)1\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1} though. This must be provided as an input, e.g., by solving the Wetterich equation appearing at zeroth order in εn\varepsilon_{n}. In the sequel, we will focus on composite operators which could also appear within Γk[h,C¯,C;g¯]\Gamma_{k}[h,\bar{C},C;\bar{g}].

The effective average action lives on the gravitational theory space 𝒯{\mathcal{T}}. By definition, this linear space comprises all action functionals that can be constructed from the field content and are invariant under background diffeomorphism symmetry. Introducing a basis {𝒪n}\{{\mathcal{O}}_{n}\} on 𝒯{\mathcal{T}}, we can expand,333Two comments are in order. Firstly, one can obviously use different bases, say, {𝒪n}\{{\mathcal{O}}_{n}\} and {𝒪n}\{{\mathcal{O}}^{\prime}_{n}\}, to expand Γk\Gamma_{k} and 𝒪~m\tilde{\mathcal{O}}_{m}, respectively. In this work, however, we are only interested in the case 𝒪n=𝒪n{\mathcal{O}}_{n}={\mathcal{O}}_{n}^{\prime}. Secondly, the expansions (8) put a non-trivial constraint towards the basis {𝒪n}\{{\mathcal{O}}_{n}\}, since our definition of Γk\Gamma_{k} includes the gauge-fixing action which also depends on the theory’s couplings. While one might separate the gauge-fixing action from Γk\Gamma_{k} in order to be able to choose the basis {𝒪n}\{{\mathcal{O}}_{n}\} more freely, here, we refrain from doing so for the sake of convenience. Moreover, in Sec. 3 below we will employ the physical gauge-fixing condition in which the gauge degrees of freedom do not source the anomalous scaling dimension of O~n\tilde{O}_{n}.

Γk[h;g¯]=nu¯n(k)𝒪n,𝒪~m(k)=nZ¯m(k)n𝒪n.\Gamma_{k}[h;\bar{g}]=\sum_{n}\,\bar{u}^{n}(k)\,{\mathcal{O}}_{n}\,,\qquad\tilde{{\mathcal{O}}}_{m}(k)=\sum_{n}\bar{Z}_{m}{}^{n}(k)\,{\mathcal{O}}_{n}\,. (8)

Here, u¯n(k)\bar{u}^{n}(k) and Z¯m(k)n\bar{Z}_{m}{}^{n}(k) are kk-dependent, dimensionful couplings. Denoting the mass-dimension of the operators as [𝒪n]=dn[{\mathcal{O}}_{n}]=-d_{n}, it is convenient to introduce their dimensionless counterparts as

un(k)u¯n(k)kdn,γm(k)nkdmdn[Z¯1tZ¯]mn.u^{n}(k)\equiv\bar{u}^{n}(k)k^{-d_{n}}\,,\qquad{\gamma}_{m}{}^{n}(k)\equiv k^{d_{m}-d_{n}}{\left[\bar{Z}^{-1}{\partial_{t}}\bar{Z}\right]_{m}}^{n}\,. (9)

The couplings unu^{n} provide natural coordinates on the theory space, while γm(k)n{\gamma}_{m}{}^{n}(k) encodes the anomalous scaling dimensions of the composite operators Pagani:2016pad .

Plugging the expansion of Γk\Gamma_{k}, Eq. (8), into the Wetterich equation (6), matching the coefficients of the basis elements 𝒪n{\mathcal{O}}_{n}, converting to dimensionless couplings (9), and solving the resulting system of equations for tun(k)\partial_{t}u^{n}(k), we arrive at the beta functions governing the dependence of the couplings un(k)u^{n}(k) on the coarse graining scale,

tun(k)=βn(u).\partial_{t}u^{n}(k)=\beta^{n}(u)\,. (10)

Analogously, the subsitution of (8) into the COE (7) yields Pagani:2016pad ; Pagani:2016dof ; Becker:2018quq ; Becker:2019tlf ; Becker:2019fhi ; Houthoff:2020zqy ; Kurov:2020csd ,

mγnmkdm𝒪m=12kdnSTr[(Γk(2)+k)1𝒪n(2)(Γk(2)+k)1ddtk],\sum_{m}{\gamma_{n}}^{m}\,k^{d_{m}}{\mathcal{O}}_{m}=-\frac{1}{2}\,k^{d_{n}}\,\text{STr}\left[\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\,{\mathcal{O}}_{n}^{(2)}\,\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\,\frac{\mathrm{d}}{\mathrm{d}t}\mathcal{R}_{k}\right]\,, (11)

In this way, γnm\gamma_{n}{}^{m} becomes a function of the couplings un(k)u^{n}(k), i.e., γn(k)mγn(u(k))m\gamma_{n}{}^{m}(k)\equiv\gamma_{n}{}^{m}(u(k)). Note that γnm\gamma_{n}{}^{m} also entails an dependence on the beta functions βn(u(k))\beta^{n}(u(k)), due to the presence of dk/dt\mathrm{d}\mathcal{R}_{k}/\mathrm{d}t under the trace.

The goal of this work is to understand the predictive power of the Reuter fixed point based on abstract properties of two-point correlation functions. To do so, we compare the stability matrix BB_{*} as in eq. (2) with scaling dimensions obtained from a composite-operator equation. Generally, one may interpret the eigenvalues of the matrix dnδnm+γnm-d_{n}\delta_{n}^{m}+\gamma_{n}{}^{m} as the full scaling dimensions of the operators {𝒪n}\{{\mathcal{O}}_{n}\}. Thus, we have two different methods at hand to evaluate critical exponents – on the one hand, the (negative) eigenvalues of the stability matrix (BFRGE)nm(B_{*}^{\rm FRGE})^{n}{}_{m} obtained from solving the Wetterich equation, and, on the other hand, the eigenvalues of (BCOE)n=mdnδnm+γnm(B_{*}^{\rm COE})_{n}{}^{m}=-d_{n}\delta_{n}^{m}+\gamma_{n}{}^{m} obtained from the composite operator equation.

Let us discuss the structure of the beta functions and the stability matrix in more detail. A critical element in their construction is the regulator k{\mathcal{R}}_{k} which is conveniently written as444More precisely, in practice one will set k(u;z)=i𝒵(i)(u¯)[Pki(z)zi]{\mathcal{R}}_{k}(u;z)=\sum_{i}{\mathcal{Z}}^{(i)}(\bar{u})\left[P_{k}^{i}(z)-z^{i}\right], where Pk(z)=z+Rk(z)P_{k}(z)=z+R_{k}(z) and the index ii runs from 1 to highest power of Δ=D¯2\Delta=-\bar{D}^{2} appearing in Γk(2)\Gamma_{k}^{(2)}. For the sake of simplicity and without loss of generality, we will remain with the simplified form (12).

k(u¯;z)=𝒵(u¯)Rk(z).{\mathcal{R}}_{k}(\bar{u};z)={\mathcal{Z}}(\bar{u})\,R_{k}(z)\,. (12)

Here, 𝒵(u¯){\mathcal{Z}}(\bar{u}) has a status of a wave function renormalization and carries a tensor structure with respect to internal field indices. It depends on tt through the dimensionful couplings u¯\bar{u} only. Furthermore, RkR_{k} is the scalar regulator. At the level of the Wetterich equation this u¯\bar{u}-dependence is dictated by the choice of regularization procedure and is a generic feature Codello:2008vh . At the level of the composite operator equation, there is more freedom since there is no direct relation between Γk(2)\Gamma_{k}^{(2)} and 𝒪n{\mathcal{O}}_{n} which would fix such a dependence based on the composite operators under consideration. As a result of the decomposition (12), one has

ddtk(u¯;z)=(n𝒵(u¯)u¯ntu¯n)Rk(z)+𝒵(u¯)(tRk).\frac{\mathrm{d}}{\mathrm{d}t}{\mathcal{R}}_{k}(\bar{u};z)=\left(\sum_{n}\frac{\partial{\mathcal{Z}}(\bar{u})}{\partial\bar{u}^{n}}\frac{\partial}{\partial t}\bar{u}^{n}\right)R_{k}(z)+{\mathcal{Z}}(\bar{u})\left(\frac{\partial}{\partial t}R_{k}\right)\,. (13)

Substituting Eqs. (8) and (13) into the Wetterich equation, equating coefficients for the basis element 𝒪n{\mathcal{O}}_{n}, and retaining the LHS==RHS form yields

tun+dnun=ωn(u)+mω~n(u)m(tum+dmum)ω^n(u;tu).\partial_{t}u^{n}+d_{n}u^{n}=\omega^{n}(u)+\sum_{m}\tilde{\omega}^{n}{}_{m}(u)\left(\partial_{t}u^{m}+d_{m}u^{m}\right)\equiv\hat{\omega}^{n}(u;\partial_{t}u)\,. (14)

This structure disentangles the classical contributions (LHS) and the contributions of the quantum fluctuations (RHS) to the flow. We may think of the latter as a vector field ω^n(u,tu)\hat{\omega}^{n}(u,\partial_{t}u) on the “theory’s configuation space”. The vector ωn(u)\omega^{n}(u) and matrix ω~n(u)m\tilde{\omega}^{n}{}_{m}(u) depend on the dimensionless couplings unu^{n}, but not on their derivatives tun\partial_{t}u^{n}. We also observe that ω~n(u)m\tilde{\omega}^{n}{}_{m}(u) may not depend on all couplings, but only those which appear in the regulator k(u¯){\mathcal{R}}_{k}(\bar{u}). Explicit expressions for ωn(u)\omega^{n}(u) and ω~n(u)m\tilde{\omega}^{n}{}_{m}(u) can be given in terms of functional traces

ωn(u)=12kdnSTr[(Γk(2)+k)1𝒵(u¯i)(tRk)]|𝒪n,ω~n(u)m=12kdmdnSTr[(Γk(2)+k)1𝒵(u¯i)u¯mRk]|𝒪n.\begin{split}\omega^{n}(u)=&\left.\frac{1}{2}\,k^{-d_{n}}\,\text{STr}\left[\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\,{\mathcal{Z}}(\bar{u}^{i})\left(\frac{\partial}{\partial t}R_{k}\right)\right]\right|_{{\mathcal{O}}_{n}}\,,\\ \tilde{\omega}^{n}{}_{m}(u)=&\left.\frac{1}{2}\,k^{d_{m}-d_{n}}\,\text{STr}\left[\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\,\frac{\partial{\mathcal{Z}}(\bar{u}^{i})}{\partial\bar{u}^{m}}\,R_{k}\right]\right|_{{\mathcal{O}}_{n}}\,.\end{split} (15)

Here, we used the notation |𝒪n|_{{\mathcal{O}}_{n}} to denote the projection of the trace onto the basis element 𝒪n{\mathcal{O}}_{n} and included factors of kk in such a way that the LHS is given by dimensionless quantities. The beta functions (10) are then obtained by solving (14) for tun\partial_{t}u^{n}:

βn(u)=dnun+m[(𝟙ω~(u))1]nωmm(u).\beta^{n}(u)=-d_{n}u^{n}+\sum_{m}\left[(\mathds{1}-\tilde{\omega}(u))^{-1}\right]^{n}{}_{m}\,\omega^{m}(u)\,. (16)

The beta functions (16) solve Eq. (14), i.e., they fulfill the identity

βn(u)=dnun+ω^n(u,β(u)).\beta^{n}(u)=-d_{n}u^{n}+\hat{\omega}^{n}(u,\beta(u))\,. (17)

We can use this identity to organize the stability matrix obtained from Eq. (16). Thus, let us consider the (generalized) stability matrix

Bn(u)m=umβn(u)B^{n}{}_{m}(u)=\frac{\partial}{\partial u^{m}}\beta^{n}(u) (18)

at a generic point along the RG trajectory,

Bn(u)m=Dn+mumω^n(u;β(u))=Dn+mumω^0n(u)+l[(umω~n(u)l)βl(u)+ω~n(u)lBl(u)m].\begin{split}B^{n}{}_{m}(u)&=-D^{n}{}_{m}+\frac{\partial}{\partial u^{m}}\hat{\omega}^{n}(u;\beta(u))\\ &=-D^{n}{}_{m}+\frac{\partial}{\partial u^{m}}\hat{\omega}_{0}^{n}(u)+\sum_{l}\left[\left(\frac{\partial}{\partial u^{m}}\tilde{\omega}^{n}{}_{l}(u)\right)\beta^{l}(u)+\tilde{\omega}^{n}{}_{l}(u)B^{l}{}_{m}(u)\right]\,.\end{split} (19)

Here, we have introduced the matrix of classical scaling dimensions, DmndnδmnD^{n}_{m}\equiv d_{n}\delta^{n}_{m}, and the vector ω^0n(u)ω^n(u;0)\hat{\omega}^{n}_{0}(u)\equiv\hat{\omega}^{n}(u;0). While it might be tempting to interpret the RHS of Eq. (19) as the classical scaling, in form of Dmn-D^{n}_{m}, and its appertaining quantum corrections, in form of mω^n\partial_{m}\hat{\omega}^{n}, one must take care with such interpretations at this stage, since mω^n\partial_{m}\hat{\omega}^{n} depends on the classical scaling dimensions dnd_{n} and the stability matrix BB itself. Solving Eq. (19) for BB, one obviously has

Bn(u)m=l[(𝟙ω~(u))1]n(Dml+ω^0l(u)um+lω~l(u)lumβl(u))l.B^{n}{}_{m}(u)=\sum_{l}\left[(\mathds{1}-\tilde{\omega}(u))^{-1}\right]^{n}{}_{l}\,\left(-D^{l}_{m}+\frac{\partial\hat{\omega}_{0}^{l}(u)}{\partial u^{m}}+\sum_{l^{\prime}}\frac{\partial\tilde{\omega}^{l}{}_{l^{\prime}}(u)}{\partial u^{m}}\,\beta^{l^{\prime}}(u)\right)\,. (20)

In order to relate this expression for the stability matrix with the anomalous scaling dimensions obtained within the composite-operator formalism, let us re-arrange Eq. (19) one more time. Having in mind that ω^n=kdn(RHS of Eq. (6)|𝒪n)\hat{\omega}^{n}=k^{-d_{n}}(\text{RHS of Eq.~{}\eqref{FRGE}}|_{{\mathcal{O}}_{n}}), it is a straightforward calculation to see that

umω^n(u;β(u))=γm(u)n+γ~m(u)n+lω~n(u)lBl(u)m,\frac{\partial}{\partial u^{m}}\hat{\omega}^{n}(u;\beta(u))=\gamma_{m}{}^{n}(u)+\tilde{\gamma}_{m}{}^{n}(u)+\sum_{l}\tilde{\omega}^{n}{}_{l}(u)B^{l}{}_{m}(u)\,, (21)

where

γm(u)n12kdmdnSTr[(Γk(2)+k)1𝒪m(2)(Γk(2)+k)1ddtk(u¯(k))]|𝒪n,\gamma_{m}{}^{n}(u)\equiv-\frac{1}{2}\,k^{d_{m}-d_{n}}\,{\rm STr}\left[\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}{{\mathcal{O}}}_{m}^{(2)}\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\,\left.\frac{\mathrm{d}}{\mathrm{d}t}{\mathcal{R}}_{k}(\bar{u}(k))\right]\right|_{{\mathcal{O}}_{n}}\,, (22)

is the anomalous-dimension matrix of Eq. (11) for the family of operators {𝒪n}\{{\mathcal{O}}_{n}\},555Note that Eq. (22) involves the relation 𝒪m(2)=Γk(2)u¯m{\mathcal{O}}_{m}^{(2)}=\frac{\partial\Gamma_{k}^{(2)}}{\partial\bar{u}^{m}}. As often the composite-operator equation (22) is evaluated for the case where {𝒪n}\{{\mathcal{O}}_{n}\} does not include gauge-fixing contributions, but where Γk\Gamma_{k} does, we will denote γ\gamma as γCOE+GF\gamma^{\mathrm{COE+GF}} and/or γCOE\gamma^{\mathrm{COE}} to distinguish the results in which the gauge-fixing term is included into {𝒪n}\{{\mathcal{O}}_{n}\} or not. and the matrix γ~m(u)n\tilde{\gamma}_{m}{}^{n}(u) contains terms involving the u¯\bar{u}-derivative of the regulator k{\mathcal{R}}_{k},

γ~m(u)n12kdmdnSTr[(Γk(2)+k)1(𝒵(u¯)u¯mRk)(Γk(2)+k)1ddtk(u¯(k))]|𝒪n+12kdmdnSTr[(Γk(2)+k)1u¯m(l𝒵(u¯)u¯ldlu¯lRk+𝒵(u¯)tRk)]|𝒪n.\begin{split}\tilde{\gamma}_{m}{}^{n}(u)\equiv&-\frac{1}{2}\,k^{d_{m}-d_{n}}\,{\rm STr}\left[\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\left(\frac{\partial{\mathcal{Z}}(\bar{u})}{\partial\bar{u}^{m}}R_{k}\right)\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\,\left.\frac{\mathrm{d}}{\mathrm{d}t}{\mathcal{R}}_{k}(\bar{u}(k))\right]\right|_{{\mathcal{O}}_{n}}\\ &+\frac{1}{2}\,k^{d_{m}-d_{n}}\,{\rm STr}\left[\left(\Gamma^{(2)}_{k}+\mathcal{R}_{k}\right)^{-1}\,\left.\frac{\partial}{\partial\bar{u}^{m}}\left(\sum_{l}\frac{\partial{\mathcal{Z}}(\bar{u})}{\partial\bar{u}^{l}}d_{l}\bar{u}^{l}R_{k}+{\mathcal{Z}}(\bar{u})\partial_{t}R_{k}\right)\right]\right|_{{\mathcal{O}}_{n}}\,.\end{split} (23)

In summary, for the stability matrix at the non-Gaussian fixed point (NGFP) we have established the formulae

(B)nm=l[(𝟙ω~(u))1]n(Dml+ω^0l(u)um|u=u)l=l[(𝟙ω~(u))1]n(Dml+ωl(u)um+um(lω~ldllul))l|u=u=l[(𝟙ω~(u))1]n(Dml+γm(u)n+γ~m(u)n)l.\begin{split}(B_{*})^{n}{}_{m}&=\sum_{l}\left[(\mathds{1}-\tilde{\omega}(u_{*}))^{-1}\right]^{n}{}_{l}\,\left(-D^{l}_{m}+\left.\frac{\partial\hat{\omega}_{0}^{l}(u)}{\partial u^{m}}\right|_{u=u_{*}}\right)\\ &=\sum_{l}\left[(\mathds{1}-\tilde{\omega}(u_{*}))^{-1}\right]^{n}{}_{l}\,\left.\left(-D^{l}_{m}+\frac{\partial\omega^{l}(u)}{\partial u^{m}}+\frac{\partial}{\partial u^{m}}\left(\sum_{l^{\prime}}\tilde{\omega}^{l}{}_{l^{\prime}}d_{l^{\prime}}u^{l^{\prime}}\right)\right)\right|_{u=u_{*}}\\ &=\sum_{l}\left[(\mathds{1}-\tilde{\omega}(u_{*}))^{-1}\right]^{n}{}_{l}\,\left(-D^{l}_{m}+\gamma_{m}{}^{n}(u_{*})+\tilde{\gamma}_{m}{}^{n}(u_{*})\right)\,.\end{split} (24)

Note that there is no simple identity expressing γm,nγ~mn\gamma_{m}{}^{n},\tilde{\gamma}_{m}{}^{n} in terms of ωn\omega^{n} and ω~nm\tilde{\omega}^{n}{}_{m}, individually. The second line of Eq. (24) should then be seen as a rearrangement of the first line, in terms of the piece given by γmn\gamma_{m}{}^{n}, appearing in the COE, and contributions associated with the dependence of the regulator on the couplings u¯\bar{u}, which have been collected in γ~mn\tilde{\gamma}_{m}{}^{n}. Only their sum is related to ω^0nm\hat{\omega}_{0}^{n}{}_{m}, namely at the fixed point via ω0^n/um(u)=γm(u)n+γ~m(u)n\partial\hat{\omega_{0}}^{n}{}/\partial u^{m}(u_{*})=\gamma_{m}{}^{n}(u_{*})+\tilde{\gamma}_{m}{}^{n}(u_{*}).

The formulae (24) are one of the main results of the present work. We stress that if the employed regulator k{\mathcal{R}}_{k} in the Wetterich equation (6) was u¯\bar{u}-independent (which in Morris:2022btf is called a non-adaptive cutoff), one has ω~=0=γ~\tilde{\omega}=0=\tilde{\gamma} and the critical exponents of the theory would be given by (B)n=mDmn+γm(u)n(B_{*})^{n}{}_{m}=-D^{n}_{m}+\gamma_{m}{}^{n}(u_{*}). For this reasoning, one might naively expect this formula to give a generally “good” approximation of the stability matrix,

(B)nmDmn+γm(u)n,(B_{*})^{n}{}_{m}\approx-D^{n}_{m}+\gamma_{m}{}^{n}(u_{*})\,, (25)

even in the case in which the regulator is coupling-dependent. In particular, one would expect that “almost-Gaussian scaling” of the stability coefficients Falls:2013bv ; Falls:2014tra ; Falls:2017lst ; Falls:2018ylp means that BB_{*} is dominated by Dmn-D^{n}_{m} with γm(u)n\gamma_{m}{}^{n}(u_{*}) providing a small correction. With the examples below in Sec. 3, we will illustrate that this is not the case though. In other words, the matrices γ~m(u)n\tilde{\gamma}_{m}{}^{n}(u_{*}) and especially ω~n(u)m\tilde{\omega}^{n}{}_{m}(u_{*}) are crucial for determining the spectrum of the stability matrix BB_{*}. This observation is important because in many systems, the critical exponents obtained from BB_{*} constitute physical observables. (Whether this is actually the case for gravity is nevertheless up to debate, due to the complexity of constructing observables.) In non-gravitational settings, it is often customary to calculate the stability matrix for fixed values of the anomalous dimension, cf. Hellwig:2015woa , following Eq. (25). This raises the question about the physical relevance of the critical exponents obtained from (B)nm(B_{*})^{n}{}_{m} versus those obtained from Dmn+γm(u)n-D^{n}_{m}+\gamma_{m}{}^{n}(u_{*}), in case there is no qualitative agreement among the two.

Lastly, we point out that for a fixed point corresponding to an asymptotically free theory, the quantum corrections to the stability matrix vanish (γ=γ~=ω~=0\gamma=\tilde{\gamma}=\tilde{\omega}=0). In this case the stability coefficients agree with the mass dimensions of the couplings.

3 Almost-Gaussian scaling within f(R)f(R)-computations

In this section, we will apply the general results obtained in Sec. 2 to the well-studied f(R)f(R)-truncations Codello:2007bd ; Machado:2007ea ; Codello:2008vh ; Benedetti:2012dx ; Dietz:2012ic ; Dietz:2013sba ; Bridle:2013sra ; Demmel:2014sga ; Demmel:2014hla ; Demmel:2015oqa ; Falls:2013bv ; Falls:2014tra ; Falls:2016wsa ; Alkofer:2018fxj ; Alkofer:2018baq ; Ohta:2015efa ; Ohta:2015fcu ; Falls:2016msz ; Mitchell:2021qjr ; Eichhorn:2015bna ; Gonzalez-Martin:2017gza ; DeBrito:2018hur ; Kluth:2020bdv . Our analysis uses the standard computational framework employed in these works: all computations are carried out using the background of a maximally symmetric dd-sphere SdS^{d}. Moreover, we employ the background approximation where the beta functions and anomalous dimensions are read off from the operators

𝒪¯n[g¯]=ddxg¯R¯n,\bar{{\mathcal{O}}}_{n}[\bar{g}]=\int\mathrm{d}^{d}x\sqrt{\bar{g}}\bar{R}^{n}\,, (26)

i.e., at zeroth order in the fluctuation fields. The scalar part of the regulator is chosen to be of Litim-type Litim:2001up ,

Rk(z)=(k2z)θ(k2z),R_{k}(z)=(k^{2}-z)\theta(k^{2}-z)\,, (27)

where Θ(x)\Theta(x) is the Heaviside step function. We first start with a one-dimensional projection in Sec. 3.1 before generalizing to the polynomial f(R)f(R)-case in Sec. 3.2.

3.1 Warm-up: the stability of Newton’s coupling

We start by illustrating the general structures derived in the previous section based on the background Einstein-Hilbert truncation Reuter:1996cp ; Reuter:2001ag ; Lauscher:2001ya ; Litim:2003vp ; Gies:2015tca , adapting the derivation given in Reuter:2019byg . Thus, we solve the Wetterich equation (6) and the COE (11) based on the geometric operators

𝒪1COE=ddxgR,𝒪1COE+GF=ddxgR+12αddxg¯g¯μνFμ[h;g¯]Fν[h;g¯].\begin{split}{\mathcal{O}}_{1}^{\rm COE}=&\int\mathrm{d}^{d}x\sqrt{g}R\,,\\ {\mathcal{O}}_{1}^{\rm COE+GF}=&\int\mathrm{d}^{d}x\sqrt{g}R+\frac{1}{2\alpha}\int\mathrm{d}^{d}x\sqrt{\bar{g}}\bar{g}^{\mu\nu}F_{\mu}[h;\bar{g}]F_{\nu}[h;\bar{g}]\,.\end{split} (28)

These operators coincide with (26) once the fluctuation fields are set to zero. The gauge-condition is taken of the general form

Fμ[h;g¯]=D¯νhμν1+βdD¯μhνν.F_{\mu}[h;\bar{g}]=\bar{D}^{\nu}h_{\mu\nu}-\frac{1+\beta}{d}\bar{D}_{\mu}h_{\nu}^{\nu}\,. (29)

This subsection chooses harmonic gauge, α=1,β=(d2)/2\alpha=1,\beta=(d-2)/2. The inclusion of the gauge-fixing term ensures that 𝒪1(2){\mathcal{O}}^{(2)}_{1} depends on the background covariant derivative D¯μ\bar{D}_{\mu} in the form of the background Laplacian Δ=g¯μνD¯μD¯ν\Delta=-\bar{g}^{\mu\nu}\bar{D}_{\mu}\bar{D}_{\nu} only.

In the first step, we project the Wetterich equation onto the subspace spanned by 𝒪¯1[g¯]\bar{{\mathcal{O}}}_{1}[\bar{g}]. Thus, all vectors and matrices associated with this example are one-dimensional. Including the hamonic gauge-fixing condition, the corresponding ansatz for the effective average action reads

Γ¯k[g]+Γ^k[h;g¯]u¯1𝒪1COE+GF,withu1k2du¯1.\bar{\Gamma}_{k}[g]+\hat{\Gamma}_{k}[h;\bar{g}]\simeq\bar{u}^{1}\,{\mathcal{O}}_{1}^{\rm COE+GF}\,,\quad\text{with}\quad u^{1}\equiv k^{2-d}\,\bar{u}^{1}\,. (30)

The coupling u¯1\bar{u}^{1} is related to Newton’s coupling GkG_{k} by u¯1=116πGk\bar{u}^{1}=-\frac{1}{16\pi G_{k}} and we have [u¯1]Du=d2>0[\bar{u}^{1}]\equiv D_{u}=d-2>0. Following the computation Reuter:2019byg , the LHS==RHS form of the projected Wetterich equation, introduced in (14), is

tu1+Duu1=ω+ω~(u1)(tu1+Duu1).\partial_{t}u^{1}+D_{u}\,u^{1}=\omega+\tilde{\omega}(u^{1})\,\left(\partial_{t}u^{1}+D_{u}\,u^{1}\right)\,. (31)

The vector ω(u)\omega(u) and matrix ω~(u)\tilde{\omega}(u) given by ω=B1\omega=B_{1} and ω~(u1)=B2/u1\tilde{\omega}(u^{1})=B_{2}/u^{1}, with the constants

B1=1(4π)d/2112(d(d3)Φd/2116(d2d+4)Φd/22),B2=1(4π)d/2124(d(d+1)Φ~d/2116d(d1)Φ~d/22).\begin{split}B_{1}=&\frac{1}{(4\pi)^{d/2}}\,\frac{1}{12}\left(d(d-3)\Phi^{1}_{d/2-1}-6(d^{2}-d+4)\Phi^{2}_{d/2}\right)\,,\\ B_{2}=&\frac{1}{(4\pi)^{d/2}}\,\frac{1}{24}\,\left(d(d+1)\tilde{\Phi}^{1}_{d/2-1}-6d(d-1)\tilde{\Phi}^{2}_{d/2}\right)\,.\end{split} (32)

The dimensionless threshold functions Φnp(w)\Phi^{p}_{n}(w) and Φ~np(w)\tilde{\Phi}^{p}_{n}(w) introduced in Reuter:1996cp are evaluated at vanishing argument and carry the dependence of the result on the choice of scalar regulator. For the Litim-type profile (27) they evaluate to

Φnp=1Γ(n+1),Φ~np=1Γ(n+2).\Phi^{p}_{n}=\frac{1}{\Gamma(n+1)}\,,\qquad\tilde{\Phi}^{p}_{n}=\frac{1}{\Gamma(n+2)}\,. (33)

Notably, the present case is special in the sense that ω\omega is independent of the coupling u1u^{1}.

Solving (31) for tu1\partial_{t}u^{1} gives the beta function

βu(u1)=Duu1+ω1ω~(u1).\beta_{u}(u^{1})=-D_{u}\,u^{1}+\frac{\omega}{1-\tilde{\omega}(u^{1})}\,. (34)

This beta function admits a single NGFP characterized by

u1=B2+B1Du,BFRGE=Du+Du2B2B1.u^{1}_{*}=B_{2}+\frac{B_{1}}{D_{u}}\,,\qquad B_{*}^{\rm FRGE}=\,-D_{u}+D_{u}^{2}\,\frac{B_{2}}{B_{1}}\,. (35)

Evaluating these expressions for d=4d=4 yields

u10.012,BFRGE=2.09.u^{1}_{*}\simeq-0.012\,,\qquad B_{*}^{\rm FRGE}=-2.09\,. (36)

Thus, we encounter a NGFP situated at a positive Newton’s coupling and a UV-attractive eigendirection.

In the next step, we compare this result to the one found in the composite operator approximation (25). Thus, we seek to find the anomalous scaling dimension of the operator 𝒪¯1[g¯]\bar{{\mathcal{O}}}_{1}[\bar{g}]. Without the inclusion of the gauge-fixing term in (28), this computation has been carried out in Houthoff:2020zqy , yielding

γCOE(u1)=1(4π)d/21u1[124(d3+9d222d24)(Φd/2212(2d)Φ~d/22)12(d33d2+4d8)(Φd/2+1312(2d)Φ~d/2+13)].\begin{split}\gamma^{\rm COE}(u^{1}_{*})=-\frac{1}{(4\pi)^{d/2}}&\,\frac{1}{u^{1}_{*}}\,\Bigg{[}\frac{1}{24}\left(d^{3}+9d^{2}-22d-24\right)\,\left(\Phi^{2}_{d/2}-\frac{1}{2}(2-d)\tilde{\Phi}^{2}_{d/2}\right)\\ &-\frac{1}{2}\left(d^{3}-3d^{2}+4d-8\right)\left(\Phi^{3}_{d/2+1}-\frac{1}{2}(2-d)\tilde{\Phi}^{3}_{d/2+1}\right)\Bigg{]}\,.\end{split} (37)

In comparison to the “exact definition” of γ\gamma, Eq. (22), this result neglects the contribution of the gauge-fixing terms in Γk(2)u¯m\frac{\partial\Gamma_{k}^{(2)}}{\partial\bar{u}^{m}}. Adding the gauge-fixing term (implementing harmonic gauge) changes this result to

γCOE+GF(u1)=1(4π)d/21u1[124d(d2+11d14)(Φd/2212(2d)Φ~d/22)12d2(d1)(Φd/2+1312(2d)Φ~d/2+13)].\begin{split}\gamma^{\rm COE+GF}(u^{1}_{*})=-\frac{1}{(4\pi)^{d/2}}&\,\frac{1}{u^{1}_{*}}\,\Bigg{[}\frac{1}{24}d\,(d^{2}+11d-14)\,\left(\Phi^{2}_{d/2}-\frac{1}{2}(2-d)\tilde{\Phi}^{2}_{d/2}\right)\\ &-\frac{1}{2}d^{2}(d-1)\left(\Phi^{3}_{d/2+1}-\frac{1}{2}(2-d)\tilde{\Phi}^{3}_{d/2+1}\right)\Bigg{]}\,.\end{split} (38)

Using these results in (25) and specializing to d=4d=4 then yields

BCOE=2+γCOE(u1)=1.91,BCOE+GF=2+γCOE+GF(u1)=1.94.B_{*}^{\rm COE}=-2+\gamma^{\rm COE}(u^{1}_{*})=-1.91\,,\qquad B_{*}^{\rm COE+GF}=-2+\gamma^{\rm COE+GF}(u^{1}_{*})=-1.94\,. (39)

Hence, we again encounter a UV-attractive eigendirection. In both computations BB_{*} is governed by the classical part Du-D_{u} and receives a subleading correction from the quantum corrections. We also observe that there is a reasonable qualitative agreement between the exact projection and its approximation by the composite operator contribution. Subsequently, we will observe how this approximation becomes increasingly worse when going to increasing the order of the truncation.

3.2 Polynomial f(R)f(R)-truncations

One of the most prominent results in the context of asymptotically safe quantum gravity is the “almost-Gaussian” scaling observed when considering polynomial f(R)f(R)-truncations at sufficiently large order of the polynomials Falls:2013bv ; Falls:2014tra . This statement paraphrases that the spectrum of critical exponents {θIFRGE=λIFRGE}\{\theta_{I}^{\mathrm{FRGE}}=-\lambda_{I}^{\mathrm{FRGE}}\} of BFRGEB^{\mathrm{FRGE}}_{*} converges to the Gaussian scaling dimensions {dI}\{d_{I}\}. Hereby, the corrections to the first three classical scaling dimensions, d0=4d_{0}=4, d1=2d_{1}=2, and d2=0d_{2}=0 are such that the first two directions remain relevant, whereas the quantum corrections shift the third direction from a marginal to a relevant direction. All other directions remain irrelevant. For operators 𝒪n{\mathcal{O}}_{n} with sufficiently large value of nn this is ensured by the aforementioned convergence pattern. Therewith, the polynomial f(R)f(R)-truncation exhibits a three-dimensional UV-critical hypersurface.

In this section, we carry out a first reconnoitring study comparing the stability matrices at the fixed point obtained from the standard FRGE procedure for the f(R)f(R)-truncation and the corresponding composite operator results. The FRGE computation is based on Eq. (24), i.e.,

(BFRGE)nm=l[(𝟙ω~(u))1]n(Dml+ωl(u)um+um(lω~ldllul))l|u=u=l[(𝟙ω~(u))1]n(Dml+γm(u)n+γ~m(u)n)l.\begin{split}(B^{\mathrm{FRGE}}_{*})^{n}{}_{m}&=\sum_{l}\left[(\mathds{1}-\tilde{\omega}(u_{*}))^{-1}\right]^{n}{}_{l}\,\left.\left(-D^{l}_{m}+\frac{\partial\omega^{l}(u)}{\partial u^{m}}+\frac{\partial}{\partial u^{m}}\left(\sum_{l^{\prime}}\tilde{\omega}^{l}{}_{l^{\prime}}d_{l^{\prime}}u^{l^{\prime}}\right)\right)\right|_{u=u_{*}}\\ &=\sum_{l}\left[(\mathds{1}-\tilde{\omega}(u_{*}))^{-1}\right]^{n}{}_{l}\,\left(-D^{l}_{m}+\gamma_{m}{}^{n}(u_{*})+\tilde{\gamma}_{m}{}^{n}(u_{*})\right)\,.\end{split} (40)

The corresponding results for the scaling corrections emerging from the COE are obtained from Eq. (25), i.e.,

(BCOE)n=mDmn+γm(u)n.(B^{\mathrm{COE}}_{*})^{n}{}_{m}=-D^{n}_{m}+\gamma_{m}{}^{n}(u_{*})\,. (41)

This comparison thus boils down to the question whether we may neglect the contributions of the matrices ω~n(u)m\tilde{\omega}^{n}{}_{m}(u_{*}) and γ~m(u)n\tilde{\gamma}_{m}{}^{n}(u_{*}). If we do so for the contribution of γ~m(u)n\tilde{\gamma}_{m}{}^{n}(u_{*}), we may think of the matrix ω~n(u)m\tilde{\omega}^{n}{}_{m}(u_{*}) as translating between the two approximations.

We start by analyzing Eq. (40). For functional f(R)f(R)-truncations the ingredients in Eq. (4) take the form

Γ¯k[g]=ddxgfk(R)=ddxgkdk(ρ),Γ^[h;g¯]=12αddxg¯g¯μνFμ[h;g¯]Fν[h;g¯].\begin{split}\bar{\Gamma}_{k}[g]&=\int\mathrm{d}^{d}x\,\sqrt{g}f_{k}(R)=\int\mathrm{d}^{d}x\,\sqrt{g}k^{d}{\mathcal{F}}_{k}(\rho)\,,\\ \hat{\Gamma}[h;\bar{g}]&=\frac{1}{2\alpha}\int\mathrm{d}^{d}x\sqrt{\bar{g}}\,\bar{g}^{\mu\nu}F_{\mu}[h;\bar{g}]F_{\nu}[h;\bar{g}]\,.\end{split} (42)

Here fk(R)f_{k}(R) is a generic, kk-dependent function of the scalar curvature RR with dimensionless counterparts, k(ρ)=kdfk(R){\mathcal{F}}_{k}(\rho)=k^{-d}f_{k}(R) and ρ=k2R\rho=k^{-2}R. Supplementing (42) by the corresponding ghost action and substituting the result into the Wetterich equation results in a partial differential equation governing the functional form of k(ρ){\mathcal{F}}_{k}(\rho). There are several incarnations of this equation which differ by the treatment of the discrete modes appearing on spherical backgrounds Morris:2022btf .

The technical construction of the flow equation for f(R)f(R)-truncations implements the York decomposition York ; York2 of the fluctuation fields

hμν=hμνTT+D¯μξν+D¯νξμ+D¯μD¯νσ1dg¯μνD¯2σ+1dg¯μνh.h_{\mu\nu}=h_{\mu\nu}^{\mathrm{TT}}+\bar{D}_{\mu}\xi_{\nu}+\bar{D}_{\nu}\xi_{\mu}+\bar{D}_{\mu}\bar{D}_{\nu}\sigma-\frac{1}{d}\bar{g}_{\mu\nu}\bar{D}^{2}\sigma+\frac{1}{d}\bar{g}_{\mu\nu}h\,. (43)

Here hμνTTh_{\mu\nu}^{\mathrm{TT}} is the transverse-traceless (TT) part, ξ^μ\hat{\xi}_{\mu} is the transverse vector field of the longitudinal part of hμνh_{\mu\nu}, σ\sigma is the corresponding longitudinal scalar, and hh is the trace part. The component fields are subject to the constraints

D¯μhμνTT=0,g¯μνhμνTT=0,D¯μξμ=0,g¯μνhμνTT=h.\bar{D}^{\mu}h_{\mu\nu}^{\mathrm{TT}}=0\,,\quad\bar{g}^{\mu\nu}h_{\mu\nu}^{\mathrm{TT}}=0\,,\quad\bar{D}_{\mu}\xi^{\mu}=0\,,\quad\bar{g}^{\mu\nu}h_{\mu\nu}^{\mathrm{TT}}=h\,. (44)

The fields are re-scaled to be orthonormal on Einstein spaces,

ξ^μ=(2(ΔR¯/d))1/2ξμ,σ^=(d1dΔ(ΔR¯/(d1)))1/2σ,h^=1dh,\hat{\xi}_{\mu}=(2(\Delta-\bar{R}/d))^{1/2}\xi_{\mu}\,,\quad\hat{\sigma}=\left(\frac{d-1}{d}\Delta\left(\Delta-\bar{R}/(d-1)\right)\right)^{1/2}\sigma\,,\quad\hat{h}=\frac{1}{\sqrt{d}}h\,, (45)

which also accounts for the Jacobians arising from the decomposition (43).

Owed to the higher-derivative terms generated by Γ¯k\bar{\Gamma}_{k} it turns out to be convenient to restrict (42) to Landau-type gauges, which fix β=0\beta=0 and take the limit α0\alpha\rightarrow 0 once all ingredients in the traces are computed. The ghost-sector arising from this “physical gauge” follows Machado:2007ea . This choice leads to a number of simplifications. Firstly, the matrix (Γk(2)+k)(\Gamma_{k}^{(2)}+{\mathcal{R}}_{k}) becomes diagonal in field space. Secondly, the contributions of the unphysical transverse vector field ξμ\xi_{\mu} and the unphysical scalar field σ\sigma become independent of the function fk(R)f_{k}(R). This is readily deduced by analyzing the α\alpha-dependence of the decomposed propagator (see Table 1 in Machado:2007ea and Eqs. (97)-(99) of Codello:2008vh ). Conversely, the matrix elements for the transverse traceless and h^\hat{h}-fluctuations appearing in (Γk(2)+k)(\Gamma_{k}^{(2)}+{\mathcal{R}}_{k}) are solely governed by Γ¯k[g]\bar{\Gamma}_{k}[g]. Specifying to the Litim-type regulator, they read Codello:2008vh

(Γk,TT(2)+k,TT)=12[(k22(d2)d(d1)R¯)fk+fk] 1TT,(Γk,h^h^(2)+k,h^h^)=d24[4(d1)2d(d2)fk′′(k2R¯d1)2+2(d1)dfk(k2R¯d1)2R¯dfk+fk].\begin{split}(\Gamma_{k,\mathrm{TT}}^{(2)}+{\mathcal{R}}_{k,\mathrm{TT}})=&-\frac{1}{2}\Bigg{[}\left(k^{2}-\frac{2(d-2)}{d(d-1)}\bar{R}\right)f^{\prime}_{k}+f_{k}\Bigg{]}\,\mathds{1}_{\mathrm{TT}}\,,\\ (\Gamma_{k,\hat{h}\hat{h}}^{(2)}+{\mathcal{R}}_{k,\hat{h}\hat{h}})=&\frac{d-2}{4}\Bigg{[}\frac{4(d-1)^{2}}{d(d-2)}f^{\prime\prime}_{k}\left(k^{2}-\frac{\bar{R}}{d-1}\right)^{2}\\ &\qquad\qquad+\frac{2(d-1)}{d}f^{\prime}_{k}\left(k^{2}-\frac{\bar{R}}{d-1}\right)-\frac{2\bar{R}}{d}f^{\prime}_{k}+f_{k}\Bigg{]}\,.\end{split} (46)

Here, 𝟙TT\mathds{1}_{\mathrm{TT}} is the identity on the space of transverse-traceless fields, fk(R)=Rfk(R)f^{\prime}_{k}(R)=\partial_{R}f_{k}(R), and we adopt an identical notation for the dimensionless functions k(ρ)=ρk(ρ){\mathcal{F}}_{k}^{\prime}(\rho)=\partial_{\rho}{\mathcal{F}}_{k}(\rho). The explicit form of corresponding matrix element of the regulator k{\mathcal{R}}_{k} arising from implementing the adaptive cutoff ΔΔ+Rk\Delta\mapsto\Delta+R_{k} is

k,TT=12fkRk 1TT,k,h^h^=d12d[(d2)fk+2(d1)fk′′(2Δ+Rk)4fk′′R¯]Rk.\begin{split}{\mathcal{R}}_{k,\mathrm{TT}}=&-\frac{1}{2}\,f^{\prime}_{k}\,R_{k}\,\mathds{1}_{\mathrm{TT}}\,,\\ {\mathcal{R}}_{k,\hat{h}\hat{h}}=&\frac{d-1}{2d}\Big{[}(d-2)f^{\prime}_{k}+2(d-1)f^{\prime\prime}_{k}(2\Delta+R_{k})-4f^{\prime\prime}_{k}\bar{R}\Big{]}R_{k}\,.\end{split} (47)

The Wetterich equation takes the schematic form

ddtΓ¯k=12TrTT[(Γ¯k,TT(2)+k,TT)1ddtk,TT]+12TrS[(Γ¯k,h^h^(2)+k,h^h^)1ddtk,h^h^]+Γ¯k-independent terms,\begin{split}\frac{\mathrm{d}}{\mathrm{d}t}\bar{\Gamma}_{k}=&\,\frac{1}{2}\operatorname{Tr}_{\mathrm{TT}}\left[\left(\bar{\Gamma}_{k,\mathrm{TT}}^{(2)}+{\mathcal{R}}_{k,\mathrm{TT}}\right)^{-1}\frac{\mathrm{d}}{\mathrm{d}t}{\mathcal{R}}_{k,\mathrm{TT}}\right]+\frac{1}{2}\operatorname{Tr}_{\mathrm{S}}\left[\left(\bar{\Gamma}_{k,{\hat{h}\hat{h}}}^{(2)}+{\mathcal{R}}_{k,{\hat{h}\hat{h}}}\right)^{-1}\frac{\mathrm{d}}{\mathrm{d}t}{\mathcal{R}}_{k,{\hat{h}\hat{h}}}\right]\\ &+\bar{\Gamma}_{k}\text{-independent terms}\,,\end{split} (48)

where the Γ¯k\bar{\Gamma}_{k}-independent terms are due to the Fadeev-Popov ghosts and the contributions of the ξμ\xi_{\mu} and σ\sigma modes. Owed to the presence of the higher-derivative terms, the regulator does not obey the simple factorization (12). Nevertheless, one is able to separate the trace contributions into parts associated with the regulator and the couplings appearing within (47). The result of this computation is rather lengthy and given as Eq. (78) of Appendix B.

We then project Eq. (48) onto the basis (26). Introducing the dimensionless volume V~=kdddxg¯ρd/2\tilde{V}=k^{d}\int\mathrm{d}^{d}x\sqrt{\bar{g}}\propto\rho^{-d/2}, the LHS of Eq. (48) is

ddtΓ¯k=V~{(tk)(ρ)+dk(ρ)2ρk(ρ)}.\frac{\mathrm{d}}{\mathrm{d}t}\bar{\Gamma}_{k}=\tilde{V}\left\{(\partial_{t}{\mathcal{F}}_{k})(\rho)+d\,{\mathcal{F}}_{k}(\rho)-2\rho{\mathcal{F}}_{k}^{\prime}(\rho)\right\}\,. (49)

Since in general the regulator is of the form k=k[fk,fk′′]{\mathcal{R}}_{k}={\mathcal{R}}_{k}[f_{k}^{\prime},f_{k}^{\prime\prime}], the RHS can be written as

RHS=V~{I0[k](ρ)+(kd+2ddtfk)(ρ)I1[k](ρ)+(kd+4ddtfk′′)(ρ)I2[k](ρ)}.\text{RHS}=\tilde{V}\left\{I_{0}[{\mathcal{F}}_{k}](\rho)+\left(k^{-d+2}\frac{\mathrm{d}}{\mathrm{d}t}f_{k}^{\prime}\right)(\rho)I_{1}[{\mathcal{F}}_{k}](\rho)+\left(k^{-d+4}\frac{\mathrm{d}}{\mathrm{d}t}f_{k}^{\prime\prime}\right)(\rho)I_{2}[{\mathcal{F}}_{k}](\rho)\right\}\,. (50)

Each function I0I_{0}, I1I_{1}, and I2I_{2} contains contribution from the TT- and trace-sector as well as from the fk(R)f_{k}(R)-independent terms. For details, we refer the reader to Falls:2014tra .

The transition from the functional to the polynomial form of Eqs. (49) and (50) uses the expansion

fk(R)=n=0Npropu¯n(k)Rn,k(ρ)=n=0Npropun(k)ρn,f_{k}(R)=\sum_{n=0}^{N_{\mathrm{prop}}}\bar{u}^{n}(k)R^{n}\,,\quad{\mathcal{F}}_{k}(\rho)=\sum_{n=0}^{N_{\mathrm{prop}}}u^{n}(k)\rho^{n}\,, (51)

where the polynomials are truncated at NpropN_{\rm prop}. The LHS given in Eq. (49) then becomes

ddtΓ¯k=V~n=0Nprop{tun+dnun}ρn,\frac{\mathrm{d}}{\mathrm{d}t}\bar{\Gamma}_{k}=\tilde{V}\sum_{n=0}^{N_{\mathrm{prop}}}\left\{\partial_{t}u^{n}+d_{n}u^{n}\right\}\rho^{n}\,, (52)

with dn=d2nd_{n}=d-2n the canonical mass dimension of the dimensionful couplings u¯n\bar{u}^{n}. Projecting the RHS, Eq. (50), onto the operators (26) is akin to Taylor-expanding this RHS around ρ=0\rho=0,

I0[k](ρ)=nωn(u)ρn,I_{0}[{\mathcal{F}}_{k}](\rho)=\sum_{n}\omega^{n}(u)\rho^{n}\,, (53)

and

(kd+2ddtfk)(ρ)I1[k](ρ)+(kd+4ddtfk′′)(ρ)I2[k](ρ)=n,mω~n(u)m[tum+dmum]ρn.\left(k^{-d+2}\frac{\mathrm{d}}{\mathrm{d}t}f_{k}^{\prime}\right)(\rho)I_{1}[{\mathcal{F}}_{k}](\rho)+\left(k^{-d+4}\frac{\mathrm{d}}{\mathrm{d}t}f_{k}^{\prime\prime}\right)(\rho)I_{2}[{\mathcal{F}}_{k}](\rho)=\sum_{n,m}\tilde{\omega}^{n}{}_{m}(u)\left[\partial_{t}u^{m}+d_{m}u^{m}\right]\rho^{n}\,. (54)

Since the LHS of Eq. (54) is independent of u¯0\bar{u}^{0}, we have ω~n00\tilde{\omega}^{n}{}_{0}\equiv 0. The structure highlighted on its RHS then arises from the expansions

(kd+2ddtfk)(ρ)=n=0n[tun+dnun]ρn1,(kd+4ddtfk′′)(ρ)=n=0n(n1)[tun+dnun]ρn2.\begin{split}\left(k^{-d+2}\frac{\mathrm{d}}{\mathrm{d}t}f_{k}^{\prime}\right)(\rho)&=\sum_{n=0}n[\partial_{t}u^{n}+d_{n}u^{n}]\rho^{n-1}\,,\\ \left(k^{-d+4}\frac{\mathrm{d}}{\mathrm{d}t}f_{k}^{\prime\prime}\right)(\rho)&=\sum_{n=0}n(n-1)[\partial_{t}u^{n}+d_{n}u^{n}]\rho^{n-2}\,.\end{split} (55)

In this way, the definitions of the vector ωn(u)\omega^{n}(u) and the matrix ω~n(u)m\tilde{\omega}^{n}{}_{m}(u) agree precisely with the definitions given in Eq. (15). The explicit form for ωn(u)\omega^{n}(u) and ω~n(u)m\tilde{\omega}^{n}{}_{m}(u) are readily obtained using Computer-Algebra-Sytems.

Combining these results with Eqs. (16) and (24), it is straightforward to obtain fixed points {un}\{u^{n}_{*}\} as well as the stability matrix BFRGEB_{*}^{\rm FRGE}. In the present work, for the case d=4d=4, we will employ the particular fixed point values {un}\{u^{n}_{*}\} obtained in Codello:2008vh . For Nprop=8N_{\mathrm{prop}}=8 these are given by (see (Codello:2008vh, , Table III)):

(u0,u1,,u8)=103(5.066,20.748,0.088,8.581,8.926,6.808,1.165,6.196,4.695).(u^{0}_{*},u^{1}_{*},\dots,u^{8}_{*})=10^{-3}(5.066,-20.748,0.088,-8.581,-8.926,-6.808,1.165,6.196,4.695)\,. (56)

When using fixed point values for propagators with Nprop<8N_{\mathrm{prop}}<8, we will simply cut off the list (56) at the corresponding value instead of using the fixed point values obtained from the Nprop<8N_{\mathrm{prop}}<8-truncation. We do so because the list (56) represents a more accurate approximation of the fixed point position {u}\{u_{*}\} also at lower orders of the truncation.

In order to compare these results with the scaling corrections obtained from γmn\gamma_{m}{}^{n} for the family of composite operators {𝒪n=ddxgRn|n=0,1,,Nscal}\{{\mathcal{O}}_{n}=\int\mathrm{d}^{d}x\sqrt{g}R^{n}\ |\ n=0,1,\dots,N_{\mathrm{scal}}\}, we must evaluate Eq. (11). In this case only the TT- and trace-modes contribute,

mγn(u)mρm=:I1+I2,\begin{split}\sum_{m}\gamma_{n}{}^{m}(u)\rho^{m}=:&\,I_{1}+I_{2}\,,\end{split} (57)

where

I1:=kdn2V~TrTT[(ddtk,TT)(Γ¯k,TT(2)+k,TT)1𝒪n(2)(Γ¯k,TT(2)+k,TT)1TT],I2:=kdn2V~TrS[(ddtk,h^h^)(Γ¯k,h^h^(2)+k,h^h^)1𝒪n(2)(Γ¯k,h^h^(2)+k,h^h^)1h^h^].\begin{split}I_{1}:=&-\frac{k^{d_{n}}}{2\tilde{V}}\operatorname{Tr}_{\mathrm{TT}}\bigg{[}\left(\frac{\mathrm{d}}{\mathrm{d}t}{\mathcal{R}}_{k,\mathrm{TT}}\right)\left({\bar{\Gamma}_{k,\mathrm{TT}}}^{(2)}+{{\mathcal{R}}_{k,\mathrm{TT}}}\right)^{-1}{{\mathcal{O}}_{n}^{(2)}}{}_{\mathrm{TT}}\left({\bar{\Gamma}_{k,\mathrm{TT}}}^{(2)}+{\mathcal{R}}_{k,\mathrm{TT}}\right)^{-1}\bigg{]}\,,\\ I_{2}:=&-\frac{k^{d_{n}}}{2\tilde{V}}\operatorname{Tr}_{\mathrm{S}}\bigg{[}\left(\frac{\mathrm{d}}{\mathrm{d}t}{\mathcal{R}}_{k,\hat{h}\hat{h}}\right){}\left({\bar{\Gamma}_{k,\hat{h}\hat{h}}}^{(2)}+{\mathcal{R}}_{k,\hat{h}\hat{h}}\right)^{-1}{{\mathcal{O}}_{n}^{(2)}}{}_{\hat{h}\hat{h}}\left({\bar{\Gamma}_{k,\hat{h}\hat{h}}}^{(2)}+{\mathcal{R}}_{k,\hat{h}\hat{h}}\right)^{-1}\bigg{]}\,.\\ \end{split} (58)

In order to read off the anomalous-dimension matrix γnm\gamma_{n}{}^{m}, we have to calculate the two traces I1I_{1} and I2I_{2} that built the RHS of Eq. (57), and Taylor-expand them in powers of ρ\rho. For the propagators arising from the Einstein-Hilbert truncation, this computation has been carried out in Houthoff:2020zqy ; Kurov:2020csd . In order to be able to compare to the polynomial f(R)f(R)-truncations of the Wetterich equation, we extend this study to the propagators (46) retaining terms in the expansion (51) up to NpropN_{\rm prop}. Since this particular calculation has not been performed in the literature before, we have collected the details in Appendix A. Note that the composite operator equation (57) allows us to truncate the operator bases {𝒪n=ddxgRn,n=0,,Nscal}\{{\mathcal{O}}_{n}=\int\mathrm{d}^{d}x\sqrt{g}R^{n},n=0,\cdots,N_{\rm scal}\} and the propagator at different values. These are then denoted by NscalN_{\mathrm{scal}} and NpropN_{\mathrm{prop}}, respectively. The FRGE formalism above obviously only admits the case Nscal=NpropN_{\mathrm{scal}}=N_{\mathrm{prop}}. As for the case of the stability matrix BFRGEB_{*}^{\rm FRGE}, results for γnm\gamma_{n}{}^{m} are readily obtained using Computer-Algebra-Software.

At this point, we are ready to compare the stability matrices (40) and (41). For concreteness, we limit the discussion to d=4d=4. Inspecting Eq. (41), one might naively infer from here that the spectrum of BCOEB^{\mathrm{COE}}_{*} is quite similar to that of BFRGEB^{\mathrm{FRGE}}_{*}, with the anomalous scaling dimensions γm(u)n\gamma_{m}{}^{n}(u_{*}) providing corrections to Dmn-D^{n}_{m} that become smaller and smaller as Nprop=NscalN_{\mathrm{prop}}=N_{\mathrm{scal}}\to\infty. In Fig. 1, we illustrate the entries of their building blocks D+γT-D+\gamma^{\rm T} (top left diagram), ω~\tilde{\omega} (top right diagram) and BB_{*} (bottom) for various values of NscalN_{\mathrm{scal}} and NpropN_{\mathrm{prop}} on a logarithmic scale.

Refer to caption
Refer to caption
Refer to caption
Figure 1: Illustration of the size of the matrix entries appearing in D+γT(u)-D+\gamma^{\mathrm{T}}(u_{*}) (upper left), ω~(u)\tilde{\omega}(u_{*}) (upper right) and the stability matrix BB_{*} (lower row) entering the master equation (24) on a logarithmic scale. The colors refer to the systems with Nscal=Nprop=3N_{\rm scal}=N_{\rm prop}=3 (cyan), Nscal=Nprop=5N_{\rm scal}=N_{\rm prop}=5 (yellow), and Nscal=Nprop=7N_{\rm scal}=N_{\rm prop}=7 (blue). The top lines illustrate that the size of the entries grow with the matrix indices m,nm,n. At the level of the stability matrix, it is the interplay between the two contributions in the top row, which significantly reduces the size of the coefficients appearing in Eq. (40).

This decomposition actually establishes that the opposite is true: The entries of γm(u)n\gamma_{m}{}^{n}(u_{*}) grow exponentially as Nprop=NscalN_{\mathrm{prop}}=N_{\mathrm{scal}}\to\infty. In addition, Fig. 1 also depicts that the proliferation of matrix entries is not limited to γm(u)n\gamma_{m}{}^{n}(u_{*}), but also happens in an almost identical fashion for ω~n(u)m\tilde{\omega}^{n}{}_{m}(u_{*}). It is striking to observe how in Eq. (24) these two rampant, and seemingly unphysical, matrices combine to BFRGEB^{\mathrm{FRGE}}_{*} whose entries’ magnitudes are in perfect unison with Dmn-D^{n}_{m}.

As a consequence of the exponential growths in the quantum parts, the spectra of BCOEB^{\mathrm{COE}}_{*} and BFRGEB^{\mathrm{FRGE}}_{*} disagree even at a qualitative level, cf. Table 1. Qualitative agreement – in terms of an agreement on the sign of the critical exponents – is only assured for small-size truncations up to Nprop=Nscal=2N_{\mathrm{prop}}=N_{\mathrm{scal}}=2. For larger-size truncations, the entries of γm(u)n\gamma_{m}{}^{n}(u_{*}) reach a size that distorts the spectrum of BCOEB^{\mathrm{COE}}_{*} altogether, with the quantum corrections exceeding the classical (Gaussian) scaling dimensions by orders of magnitude. In Table 1 this becomes evidently clear for Nprop=Nscal=6N_{\mathrm{prop}}=N_{\mathrm{scal}}=6. These corrections surely seem intuitively anomalous and far from a physically reliable estimate. This is a paradoxical situation since, as we had mentioned earlier on, there exist plenty of physical systems, e.g. in the realm of condensed matter physics, for which the critical exponents emanating from BCOEB^{\mathrm{COE}}_{*} constitute actual physical observables.

As regards to observables, we would like to point out another striking results that Table 1 exhibits. For geometrical operators depending on some characteristic scale, the (negative) eigenvalues of D+γ-D+\gamma can also be interpreted as the geometric scaling dimensions of these operators. In case of a single-operator approximation given by the spacetime volume 𝒪0{\mathcal{O}}_{0} (i.e., Nscal=0N_{\mathrm{scal}}=0), we find dimensional reduction of its scaling dimension from classically 4 down to 2.042.04. Albeit a gauge-dependent result, this strikingly confirms the dimensional reduction of spacetime in the UV from 4 to 2 dimensions that has already been observed in a variety of settings Carlip:2019onx , including Asymptotic Safety Reuter:2012xf and CDT Loll:2019rdj .

NscalN_{\rm scal} NpropN_{\rm prop} λ0\lambda_{0} λ1\lambda_{1} λ2\lambda_{2} λ3\lambda_{3} λ4\lambda_{4} λ5\lambda_{5} λ6\lambda_{6}
D-D 4-4 2-2 0 22 44 66 88
0 3\geq 3 BCOEB_{*}^{\rm COE} 2.04-2.04
11 11 .BFRGE\Big{.}B_{*}^{\rm FRGE} 2.09-2.09
.BCOE\Big{.}B_{*}^{\rm COE} 1.91-1.91
.BCOE+GF\Big{.}B_{*}^{\rm COE+GF} 1.94-1.94
22 22 .BFRGE\Big{.}B_{*}^{\rm FRGE} 27.02-27.02 1.26±2.45i-1.26\pm 2.45\mathrm{i}
.BCOE\Big{.}B_{*}^{\rm COE} 4.06±1.38i-4.06\pm 1.38\mathrm{i} 0.89-0.89
44 44 .BFRGE\Big{.}B_{*}^{\rm FRGE} 2.84±2.42i-2.84\pm 2.42\mathrm{i} 1.54-1.54 4.274.27 5.095.09
.BCOE\Big{.}B_{*}^{\rm COE} 3.07±2.79i-3.07\pm 2.79\mathrm{i} 3.833.83 49.149.1 290.7290.7
66 66 .BFRGE\Big{.}B_{*}^{\rm FRGE} 2.39±2.38i-2.39\pm 2.38\mathrm{i} 1.51-1.51 4.164.16 4.68±6.09i4.68\pm 6.09\mathrm{i} 8.688.68
.BCOE\Big{.}B_{*}^{\rm COE} 21.30±22.10i-21.30\pm 22.10\mathrm{i} 4.53-4.53 0.66±2.35i0.66\pm 2.35\mathrm{i} 329.0329.0 954.9954.9
.Bloc\Big{.}B_{*}^{\rm loc} (26+4.4d1)1010-(26+4.4d_{1})10^{10} 6-6 4-4 2-2 0 22
66 22 .BFRGE\Big{.}B_{*}^{\rm FRGE} - - - - -
.BCOE\Big{.}B_{*}^{\rm COE} 11.74±19.49i-11.74\pm 19.49\mathrm{i} 9.80-9.80 2.41±1.88i-2.41\pm 1.88\mathrm{i} 0.15-0.15 3.643.64
Table 1: Illustration of the spectral properties of the matrices BFRGEB_{*}^{\rm FRGE} and BCOEB_{*}^{\rm COE} for several values of NscalN_{\rm scal} and NpropN_{\rm prop}. The first line (classical scaling) gives the entries of D-D. Complex eigenvalues are presented pairwise in a double-column entry. The case Nscal=Nprop=1N_{\rm scal}=N_{\rm prop}=1 summarizes the results of Sect. 3.1 while the lower entries compare the spectra of the stability matrix obtained from solving the FRGE and COE for polynomial f(R)f(R)-truncations. Note that the definition of BFRGEB_{*}^{\rm FRGE} entails that Nscal=NpropN_{\rm scal}=N_{\rm prop}. The results from the “localization approximation” (loc) are referenced from Eq. (65).

4 Almost-Gaussian Scaling from the Quantum Corrections

The take-away message from Fig. 1 is that the spectrum of the stability matrix is essentially determined by the interplay between the quantum contributions to Eq. (40). At this point, it is highly interesting to understand the mechanism that actually underlies our findings in Table 1. The key ingredient is the (regularized) two-point correlation function (Γk(2)+k)(\Gamma_{k}^{(2)}+{\mathcal{R}}_{k}). In general, this operator is matrix-valued in field space. A detailed inspection of this matrix shows, that it is only one specific matrix element which drives the mechanism, the one associated with the trace part h^\hat{h} appearing in the decomposition (43).

The first step is to understand the root structure of (46). For this purpose, we restrict to d=4d=4, substitute the expansions of f(R¯)f(\bar{R}) up to Nprop={3,4,5,6,7,8}N_{\rm prop}=\{3,4,5,6,7,8\}, and convert to dimensionless variables. Substituting the fixed point position (56), this leads to a polynomial of order NpropN_{\rm prop} in the dimensionless curvature ρ\rho. For Nprop=8N_{\rm prop}=8, the resulting expression is given by

(Γk,h^h^(2)+k,h^h^)(u)=k44(0.197ρ81.279ρ7+0.958ρ6+2.072ρ5+0.976ρ40.707ρ30.732ρ20.464ρ0.051).\begin{split}(\Gamma_{k,\hat{h}\hat{h}}^{(2)}+{\mathcal{R}}_{k,\hat{h}\hat{h}})(u_{*})=\frac{k^{4}}{4}&\,\big{(}0.197\rho^{8}-1.279\rho^{7}+0.958\rho^{6}+2.072\rho^{5}+0.976\rho^{4}\\ &\,\,-0.707\rho^{3}-0.732\rho^{2}-0.464\rho-0.051\big{)}\,.\end{split} (59)

The roots of these polynomials are easily found numerically and depicted in the left panel of Fig. 2. Increasing NpropN_{\rm prop} adds a new root in each step. The most important one is the one which is closest to the expansion point ρ=0\rho=0. This root appears at Nprop=3N_{\rm prop}=3 and remains stable when NpropN_{\rm prop} is increased. For Nprop=8N_{\rm prop}=8, its position is given by

ρ0Nprop=8=0.133.\rho_{0}^{N_{\rm prop}=8}=-0.133\,. (60)

The relative distance between the root found at lower orders, Δρρ0Nprop=nρ0Nprop=8ρ0Nprop=8\Delta\rho\equiv\frac{\rho_{0}^{N_{\rm prop}=n}-\rho_{0}^{N_{\rm prop}=8}}{\rho_{0}^{N_{\rm prop}=8}}, is displayed in the inset, demonstrating the stability of the root with respect to in increase of NpropN_{\rm prop}. The corresponding analysis in the transverse-traceless sector shows that in this case the root closest to the origin is at |ρh^h^Nprop=8|0.66|\rho_{\hat{h}\hat{h}}^{N_{\rm prop}=8}|\approx 0.66.

Refer to caption
Refer to caption
Figure 2: Illustration of the analytic structure of (Γk,h^h^(2)+k,h^h^)(\Gamma_{k,\hat{h}\hat{h}}^{(2)}+{\mathcal{R}}_{k,\hat{h}\hat{h}}) (left) and the coefficients appearing in the series expansion eq. (61) (right).

It turns out that it is the root (60) that dominates the expansion of the flow equations in the dimensionless Ricci scalar ρ\rho. This is seen as follows. The Wetterich and composite operator equations contain the inverse of the two-point function. For the specific gauge choice the matrix (Γk(2)+k)(\Gamma_{k}^{(2)}+{\mathcal{R}}_{k}) diagonalizes, so the inverse contains the inverse of (Γk,h^h^(2)+k,h^h^)(\Gamma_{k,\hat{h}\hat{h}}^{(2)}+{\mathcal{R}}_{k,\hat{h}\hat{h}}) which is expanded at ρ=0\rho=0. Evaluating this expansion at the fixed point uu_{*} yields

(Γk,h^h^(2)+k,h^h^)1(u)=nτn(u)ρn.(\Gamma_{k,\hat{h}\hat{h}}^{(2)}+{\mathcal{R}}_{k,\hat{h}\hat{h}})^{-1}(u_{*})=\sum_{n}\tau^{n}(u_{*})\rho^{n}\,. (61)

The expansion coefficients τn(u)\tau^{n}(u_{*}) are illustrated in the right panel of Fig. 2. Here the blue dots correspond to log|τn(u)|\log|\tau^{n}(u_{*})| while the orange plots display log|τn(u)(ρ0Nprop=8)n|\log|\tau^{n}(u_{*})(\rho_{0}^{N_{\rm prop}=8})^{n}|. Owed to the pole in the complex unit circle, the coefficients grow exponentially. This is illustrated by the blue dots. The orange points show that the growths is controlled by ρ0Nprop=8\rho_{0}^{N_{\rm prop}=8} already for low values of nn. Thus, we recover the familiar result that the growths of the expansion coefficients is controlled by the root closest to the expansion point666A simple example illustrating this effect is the expansion 1/(110x)=n10nxn{1}/({1-10x})=\sum_{n}10^{n}\,x^{n} where the expansion coefficients are determined by the pole at x=1/10x=1/10.

limn|τn+1(u)τn(u)|=1ρ0Nprop=8.\lim_{n\rightarrow\infty}\,\left|\frac{\tau^{n+1}(u_{*})}{\tau^{n}(u_{*})}\right|=\frac{1}{\rho_{0}^{N_{\rm prop}=8}}\,. (62)

As shown in the right diagram of Fig. 2, this relation is already fulfilled to very good accuracy at low values n10n\simeq 10. It is straightforward to repeat this analysis for the other matrix elements appearing in (Γk(2)+k)(\Gamma_{k}^{(2)}+{\mathcal{R}}_{k}). Since the root in the scalar sector is the one closest to the expansion point, the expansion coefficients in this sector exhibit the fastest growth and therefore determine the structure of the flow equation at large NpropN_{\rm prop}. This feature propagates into the entries of ωn(u)\omega^{n}(u_{*}) and ω~n(u)m\tilde{\omega}^{n}{}_{m}(u_{*}) and also to γm(u)n\gamma_{m}{}^{n}(u_{*}) and γ~m(u)n\tilde{\gamma}_{m}{}^{n}(u_{*}).

The exponential growth of the expansion coefficients (61) entails that it is the scalar trace which dominates the beta functions for couplings u¯n\bar{u}^{n} with nn sufficiently large. In particular, the classical contributions originating from the LHS of the flow equation (encoded in DD) are negligible. This suggests that for sufficiently large values of nn, a good approximation of the flow is given by

12TrS[(Γk,h^h^(2)+k,h^h^)1(ddtk,h^h^)]0.\frac{1}{2}{\rm Tr}_{S}\bigg{[}\left({\Gamma_{k,\hat{h}\hat{h}}}^{(2)}+{\mathcal{R}}_{k,\hat{h}\hat{h}}\right)^{-1}\,\left(\frac{\mathrm{d}}{\mathrm{d}t}{\mathcal{R}}_{k,\hat{h}\hat{h}}\right)\bigg{]}\approx 0\,. (63)

We will call this the “localization” (loc) approximation, since the analytic structure of the propagator essentially localizes the contributions of the RG flow in one single sector.

Since (63) is a pure quantum expression, this immediately raises the puzzle about how the localization approximation may recover any knowledge about the classical scaling dimensions dund_{u^{n}}. In order to answer this question, let us evaluate the RG-time derivative of (47),

ddtk,h^h^=34{[fk+6fk′′(Δ+Rk)2fk′′R¯]tRk+[(dfk/dt)+3(2Δ+Rk)(dfk′′/dt)2R¯(dfk′′/dt)]Rk},\begin{split}\frac{\mathrm{d}}{\mathrm{d}t}{\mathcal{R}}_{k,\hat{h}\hat{h}}=\frac{3}{4}&\Big{\{}\left[f^{\prime}_{k}+6f^{\prime\prime}_{k}(\Delta+R_{k})-2f^{\prime\prime}_{k}\bar{R}\right]\partial_{t}R_{k}\\ &\;+\left[(\mathrm{d}f^{\prime}_{k}/\mathrm{d}t)+3(2\Delta+R_{k})(\mathrm{d}f^{\prime\prime}_{k}/\mathrm{d}t)-2\bar{R}(\mathrm{d}f^{\prime\prime}_{k}/\mathrm{d}t)\right]R_{k}\Big{\}}\,,\end{split} (64)

where the tt-derivatives of fk(R)f_{k}(R) where given in Eq. (55). The first line in (64) will give the contributions to ωn(u)\omega^{n}(u_{*}) while the second encodes the terms appearing in ω~n(u)m\tilde{\omega}^{n}{}_{m}(u_{*}). Moreover, Eq. (55) shows that (63) actually gives beta functions for the couplings unu^{n}, 1nNprop1\leq n\leq N_{\rm prop}.

We can then evaluate the trace (63) using the zeroth order in the heat kernel and approximating the propagator by its expression at the fixed point uu_{*}. The approximation in the heat-kernel expansion is again justified by the exponential growth of the series coefficients τn\tau^{n} to which higher-order contributions of the heat-kernel are subleading. Solving for tun\partial_{t}u^{n}, this results in beta functions which are linear in the couplings unu^{n}. The stability matrix is then obtained by taking the derivative with respect to these couplings. This results in the upper-triangular matrix. Thus the eigenvalues of BlocB_{*}^{\rm loc} are given by its diagonal entries. For example, Nprop=6N_{\rm prop}=6 leads to

Bloc=((4.4×1010d12.6×1011)3.68×1016243.1458.6561.2.53×104.0d2627162729196837.00d36125414587.000d4692437.0000d56547.00000d66)B_{*}^{\rm loc}=\left(\begin{array}[]{cccccc}(-4.4\times 10^{10}d_{1}-2.6\times 10^{11})&-3.68\times 10^{16}&-243.&-1458.&-6561.&-2.53\times 10^{4}\\ \Big{.}0&-d_{2}-6&27&162&729&\frac{19683}{7}\\ \Big{.}0&0&-d_{3}-6&12&54&\frac{1458}{7}\\ \Big{.}0&0&0&-d_{4}-6&9&\frac{243}{7}\\ \Big{.}0&0&0&0&-d_{5}-6&\frac{54}{7}\\ \Big{.}0&0&0&0&0&-d_{6}-6\\ \end{array}\right) (65)

Notably, these contain the classical scaling dimension of the couplings. Thus one can indeed recover an almost classical scaling from (63). The specific structure of the regulator (47) is crucial for that, however.

This result also clarifies why the COE does not recover the almost-Gaussian scaling behavior. The localization mechanism owed to the pole in the propagator is still operational in this case. However, the regulator is no longer tied to the operators 𝒪n{\mathcal{O}}_{n}. In particular k{\mathcal{R}}_{k} is independent of the anomalous scaling dimension. Thus, the mechanism leading to stability matrices of the form (65) fails in this case.

5 Discussion and Conclusions

Interacting renormalization group (RG) fixed points provide an attractive mechanism for providing a consistent and predictive high-energy completion of a theory beyond the realm of perturbation theory. The predictive power of such a fixed point is encoded in the spectrum of the stability matrix BB_{*}, Eq. (2), controlling the linearized RG flow in its vicinity. The master equation derived in Eq. (24) encodes the general structure of this matrix separating the contributions originating from the classical scaling, quantum corrections, and regulator piece. This result is general in the sense that it holds for any fixed point irrespective of a given field content. Results obtained from the Wetterich equation and the composite operator equation can readily be cast into this parameterization. This highlights the structural differences occurring when evaluating the operator spectrum within these distinct computation schemes.

We then applied the general formalism to the Reuter fixed point projected on f(R)f(R)-type interactions polynomial in the Ricci scalar RR. The evaluation of the stability matrix based on the Wetterich equation yields a spectrum of BB_{*} which almost follows the canonical scaling for large powers of RR. This is a key result supporting the predictive power of the Reuter fixed point. In the composite operator equation, this feature is absent though. Our general framework provides a comprehensive understanding of this mismatch in the following way: near the Reuter fixed point the regularized two-point function associated with the trace sector in the gravitational fluctuations exhibits a pole close to the expansion point of the functional renormalization group equations. This pole dominates the structure of the stability matrix for polynomials of sufficiently high order in RR. The flow equations are well-approximated by the quantum contributions stemming from the trace-modes and the classical part of the stability matrix does not play any role. This mechanism is operative for both the Wetterich and the composite operator equation. The difference in the spectra of BB_{*} can then be traced to the different regulator prescriptions. The regularization based on the Wetterich equation is constructed in such a way that also the dominant quantum contributions to the stability matrix exhibit classical scaling properties. For the composite operator equation this feature of the regulator and hence the almost-Gaussian scaling of the spectrum of BB is absent. We stress that low-order expansions are unaffected by this mechanism though and shows qualitative agreement within the two complementary approaches.

Similar observations about the regulator dependence of the spectrum have recently been made in the context of functional f(R)f(R)-truncations using non-adaptive regularization schemes Mitchell:2021qjr ; Morris:2022btf . Studying the spectrum of admissible perturbations at a fixed functional f(R)f_{*}(R) using Sturm-Liouville theory, it is found that the spectrum of BB_{*} may either come with a finite or infinite number of relevant eigendirections, depending on the precise implementation of the regulator in the trace sector Mitchell:2021qjr .

These findings raise the crucial question whether the almost-Gaussian scaling is indeed a genuine feature of the Reuter fixed point or a build-in feature of the key equation used to study it. One way to shed light on this question would be the study of f(R)f(R)-type projections using an expansion based on an operator basis of the form {𝒪n=ddxg(RR0)n}\{{\mathcal{O}}_{n}=\int\mathrm{d}^{d}x\sqrt{g}(R-R_{0})^{n}\}, where the reference curvature R0R_{0} is chosen in such a way that there are no propagator poles within the unit circle centered at the expansion point. Alternatively, a detailed understanding of the poles contained in (Γk(2)+k)1(\Gamma_{k}^{(2)}+{\mathcal{R}}_{k})^{-1} on a generic, curved background spacetime may also give further information about the predictive power of the gravitational asymptotic safety program. This assessment also connects to the realm of functional f(R)f(R)-truncations where the pole structure of (Γk(2)+k)1(\Gamma_{k}^{(2)}+{\mathcal{R}}_{k})^{-1} also played a crucial role in obtaining isolated, global fixed functionals fk(R)f_{k}(R) by insisting that the solutions can pass through the poles of the flow equation located on the real axis Dietz:2012ic ; Demmel:2012ub .

We note that the mechanism uncovered in this article may actually not be limited to the realm of quantum gravity. The interplay between poles of a two-point function and the spectrum of the stability matrix may be realized in other settings as well. Thus, the effect of quantum-induced almost-Gaussian scaling may occur much more frequently. Clearly, a general understanding of this situation would be highly desirable also beyond the realm of f(R)f(R)-computations and we hope to come back to this in the future.

Acknowledgements

We thank Martin Reuter, Carlo Pagani, Luca Buoninfante, and Rudrajit Banerjee for interesting discussions. MB gratefully acknowledges financial support by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – project number 493330310. The work of AK on the spectrum of the stability matrix is supported by the Russian Science Foundation grant No 23-12-00051, https://rscf.ru/en/project/23-12-00051/.

Appendix A Calculation of γnm\gamma_{n}{}^{m} for the operator family {ddxgRn}\{\int\mathrm{d}^{d}x\sqrt{g}R^{n}\}

This appendix collects the technical details entering into the evaluation of the operator traces in Eq. (57). The computation utilizes the heat-kernel technology reviewed in Codello:2008vh and we outline the relevant steps in order to render the present work self-contained.

Inspecting the traces in Eq. (57), we see that their arguments are given by operator-valued functions Ws(Δ)W_{s}(\Delta) where Δ:=g¯μνD¯μD¯ν\Delta:=-\bar{g}^{\mu\nu}\bar{D}_{\mu}\bar{D}_{\nu} is the Laplacian constructed from the background metric and the trace is either over scalars (s=Ss=S) or transverse-traceless symmetric matrices (s=TT)s={\rm TT}). Our computation tracks powers of the background scalar curvature R¯\bar{R} only. Hence it suffices to work on a spherically symmetric background.

We then evaluate the traces using the early-time expansion of the heat kernel. Exploiting that we work on a background sphere, this expansion takes the form

Trs[W(Δ)]=1(4π)d/2m=0αmsQd2m[W]ddxg¯R¯m.\operatorname{Tr}_{s}\left[W(\Delta)\right]=\frac{1}{(4\pi)^{d/2}}\sum_{m=0}^{\infty}\alpha_{m}^{s}\,Q_{\frac{d}{2}-m}[W]\,\int\mathrm{d}^{d}x\sqrt{\bar{g}}\,\bar{R}^{m}\,. (66)

Here αms\alpha_{m}^{s} are the deWitt coefficients with R¯=1\bar{R}=1. For the background sphere these numbers can be found in Kluth:2019vkg and we collect the relevant ones in Table 2.

. mm 0 11 22 33 44 55 66
. αmS\alpha_{m}^{S} 11 16\frac{1}{6} 292160\frac{29}{2160} 3754432\frac{37}{54432} 1496531840\frac{149}{6531840} 179431101440\frac{179}{431101440} 1387201755473920-\frac{1387}{201755473920}
. αmTT\alpha_{m}^{\rm TT} 55 56-\frac{5}{6} 1432-\frac{1}{432} 31154432\frac{311}{54432} 1091306368\frac{109}{1306368}
Table 2: Summary of the relevant deWitt coefficients for the Laplacian on a four-dimensional background sphere Kluth:2019vkg . The coefficients not listed in the table do not enter into our computation.

The QQ-functionals are given in terms of the Mellin transform of the trace argument. For positive index m>0m>0,

Qm[W]=1Γ(m)0dzzm1W(z).Q_{m}[W]=\frac{1}{\Gamma(m)}\int_{0}^{\infty}\mathrm{d}z\,z^{m-1}W(z)\,. (67)

For m0m\leq 0 one can integrate by parts to obtain

Qm[W]=(1)kΓ(m+k)0dzzm+k1W(k)(z),m+k>0.Q_{-m}[W]=\frac{(-1)^{k}}{\Gamma(m+k)}\int_{0}^{\infty}\mathrm{d}z\,z^{m+k-1}W^{(k)}(z)\,,\quad m+k>0\,. (68)

For m+km+k being integer, this reduces to the relation Qm[W]=(1)mW(m)(0)Q_{-m}[W]=(-1)^{m}\,W^{(m)}(0). Based on Eq. (67) one furthermore establishes the identity that

Qm[zpW]=Γ(m+p)Γ(m)Qm+p[W].Q_{m}\left[z^{p}\,W\right]=\frac{\Gamma(m+p)}{\Gamma(m)}Q_{m+p}[W]\,. (69)

In order to transition to dimensionless quantities, we finally define

qmS:=(k2)m2Qm[WS(;k2ρ)],qmTT:=(k2)m1Qm[WTT(;k2ρ)].\begin{split}q_{m}^{\mathrm{S}}&:=(k^{2})^{-m-2}Q_{m}[W^{\mathrm{S}}(\,\cdot\,;k^{2}\rho)]\,,\\ q_{m}^{\mathrm{TT}}&:=(k^{2})^{-m-1}Q_{m}[W^{\mathrm{TT}}(\,\cdot\,;k^{2}\rho)]\,.\end{split} (70)

Based on these prerequisites, we proceed by evaluating I1I_{1} and I2I_{2}. The composite operator Hessian 𝒪n(2){\mathcal{O}}^{(2)}_{n} can be found in Kurov:2020csd and the propagator modalities are as in Codello:2008vh . This results in the trace arguments

WTT(Δ;R¯)=[ηfk(R¯)+t]Rk,WS(Δ;R¯)= 4(tRk+ζfk(R¯)Rk)[(d1)ΔR¯]+2(d1)Rk[2tRk+ζfk(R¯)Rk]+(d2)fk(R¯)fk′′(R¯)[tRkηfk(R¯)Rk].\begin{split}W^{\mathrm{TT}}(\Delta;\bar{R})=&\,\left[-\eta_{f^{\prime}_{k}(\bar{R})}+\partial_{t}\right]R_{k}\,,\\ W^{\mathrm{S}}(\Delta;\bar{R})=&\,4\left(\partial_{t}R_{k}+\zeta_{f_{k}(\bar{R})}R_{k}\right)\left[(d-1)\Delta-\bar{R}\right]\\ &+2(d-1)R_{k}\left[2\partial_{t}R_{k}+\zeta_{f_{k}(\bar{R})}R_{k}\right]+(d-2)\frac{f^{\prime}_{k}(\bar{R})}{f^{\prime\prime}_{k}(\bar{R})}\left[\partial_{t}R_{k}-\eta_{f_{k}(\bar{R})}R_{k}\right]\,.\end{split} (71)

Here we specialized to the Litim-type cutoff (27) and introduced the anomalous dimensions

ηfk(R)=1fk(R)tfk(R)=(d2)(tk)(ρ)k(ρ)+2ρk′′(ρ)k(ρ),ζfk(R)=1fk′′(R)tfk′′(R)=(d4)+(tk′′)(ρ)k′′(ρ)2ρk′′′(ρ)k′′(ρ).\begin{split}\eta_{f_{k}(R)}&=-\frac{1}{f^{\prime}_{k}(R)}\partial_{t}f^{\prime}_{k}(R)=-(d-2)-\frac{(\partial_{t}\mathcal{F}_{k}^{\prime})(\rho)}{\mathcal{F}_{k}^{\prime}(\rho)}+2\rho\frac{\mathcal{F}_{k}^{\prime\prime}(\rho)}{\mathcal{F}_{k}^{\prime}(\rho)}\,,\\ \zeta_{f_{k}(R)}&=\frac{1}{f^{\prime\prime}_{k}(R)}\partial_{t}f^{\prime\prime}_{k}(R)=(d-4)+\frac{(\partial_{t}\mathcal{F}_{k}^{\prime\prime})(\rho)}{\mathcal{F}_{k}^{\prime\prime}(\rho)}-2\rho\frac{\mathcal{F}_{k}^{\prime\prime\prime}(\rho)}{\mathcal{F}_{k}^{\prime\prime}(\rho)}\,.\end{split} (72)

Using the operator insertion 𝒪n=ddxgRn{\mathcal{O}}_{n}=\int\mathrm{d}^{d}x\sqrt{g}R^{n}, the two traces in Eq. (57) evaluate to

I1=1(4π)d/2k(ρ)ρn1(k(ρ)(12(d2)d(d1)ρ)+k(ρ))2×{d2c1TTα0TTqd2+1TT+m=0(c0TTραmTT+c1TT(d/2m1)αm+1TT)qd2mTT}\begin{split}I_{1}=&\frac{1}{(4\pi)^{d/2}}\mathcal{F}^{\prime}_{k}(\rho)\rho^{n-1}\Bigg{(}\mathcal{F}^{\prime}_{k}(\rho)\left(1-\frac{2(d-2)}{d(d-1)}\rho\right)+\mathcal{F}_{k}(\rho)\Bigg{)}^{-2}\\ &\times\Bigg{\{}\frac{d}{2}c_{1}^{\mathrm{TT}}\alpha_{0}^{\mathrm{TT}}q_{\frac{d}{2}+1}^{\mathrm{TT}}+\sum_{m=0}^{\infty}\bigg{(}c_{0}^{\mathrm{TT}}\rho\,\alpha_{m}^{\mathrm{TT}}+c_{1}^{\mathrm{TT}}(d/2-m-1)\alpha_{m+1}^{\mathrm{TT}}\bigg{)}q_{\frac{d}{2}-m}^{\mathrm{TT}}\Bigg{\}}\end{split} (73)

and

I2=(1d)(4π)d/2k′′(ρ)ρn2{2(d1)2k′′(ρ)[1ρd1]2+(d1)(d2)k(ρ)[1ρd1](d2)ρk(ρ)+d(d2)2k(ρ)}2×{d2[c1Sα0Sρ+(d21)c2Sα1S]qd2+1S+d2(d2+1)c2Sα0Sqd2+2S+m=0[c0SαmSρ2+(d2m1)c1Sαm+1Sρ+(d2m2)(d2m1)c2Sαm+2S]qd2mS}.\begin{split}I_{2}=\,&\frac{(1-d)}{(4\pi)^{d/2}}\,\mathcal{F}_{k}^{\prime\prime}(\rho)\rho^{n-2}\Bigg{\{}2(d-1)^{2}\mathcal{F}_{k}^{\prime\prime}(\rho)\left[1-\frac{\rho}{d-1}\right]^{2}\\ &\quad\quad\quad\quad+(d-1)(d-2)\mathcal{F}_{k}^{\prime}(\rho)\left[1-\frac{\rho}{d-1}\right]-(d-2)\rho\mathcal{F}_{k}^{\prime}(\rho)+\frac{d(d-2)}{2}\mathcal{F}_{k}(\rho)\Bigg{\}}^{-2}\\ &\times\Bigg{\{}\frac{d}{2}\left[c_{1}^{\mathrm{S}}\alpha_{0}^{\mathrm{S}}\rho+\left(\frac{d}{2}-1\right)c_{2}^{\mathrm{S}}\alpha_{1}^{\mathrm{S}}\right]q_{\frac{d}{2}+1}^{\mathrm{S}}+\frac{d}{2}\left(\frac{d}{2}+1\right)c_{2}^{\mathrm{S}}\alpha_{0}^{\mathrm{S}}q_{\frac{d}{2}+2}^{\mathrm{S}}\\ &\quad\quad+\sum_{m=0}^{\infty}\Bigg{[}c_{0}^{\mathrm{S}}\alpha_{m}^{\mathrm{S}}\rho^{2}+\left(\frac{d}{2}-m-1\right)c_{1}^{\mathrm{S}}\alpha_{m+1}^{\mathrm{S}}\rho\\ &\quad\quad\quad\quad\quad\quad+\left(\frac{d}{2}-m-2\right)\left(\frac{d}{2}-m-1\right)c_{2}^{\mathrm{S}}\alpha_{m+2}^{\mathrm{S}}\Bigg{]}q_{\frac{d}{2}-m}^{\mathrm{S}}\Bigg{\}}\,.\end{split} (74)

The coefficients csc^{s} depend on dd and nn and are given by

c0TT=n(d2)d(d1)12,c0S=d(d2)4n(dn1),c1TT=n/2,c1S=n(d2(4n1)(d1)1)2,c2S=n(n1)(d1)2.\begin{split}c_{0}^{\mathrm{TT}}=\frac{n(d-2)}{d(d-1)}-\frac{1}{2}\,,\qquad&c_{0}^{\mathrm{S}}=\frac{d(d-2)}{4}-{n(d-n-1)}\,,\\ c_{1}^{\mathrm{TT}}=-n/2\,,\qquad&c_{1}^{\mathrm{S}}=\frac{n(d^{2}-(4n-1)(d-1)-1)}{2}\,,\\ &c_{2}^{\mathrm{S}}={n(n-1)(d-1)^{2}}\,.\end{split} (75)

In principle, one can now expand Eqs. (73) and (74) in powers of ρ\rho and subsequently read off γn(u(k))m\gamma_{n}{}^{m}(u(k)) according to Eq. (57). Owed to the kk-derivatives acting on the couplings u(k)u(k) in the second term of the anomalous dimensions (72) the resulting system is not self-contained though and requires an explicit expression for the beta functions tun(k)=βn(u)\partial_{t}u^{n}(k)=\beta^{n}(u). At a fixed point uu^{*} the beta functions vanish by definition. This entails that

k(ρ)(ρ)=n=0Npropunρn,(t)(ρ)=0.\mathcal{F}_{k}(\rho)\to\mathcal{F}_{*}(\rho)=\sum_{n=0}^{N_{\mathrm{prop}}}u^{n}_{*}\rho^{n}\,,\quad(\partial_{t}\mathcal{F}_{*})(\rho)=0\,. (76)

Substituting these conditions into the expressions for the anomalous dimensions gives

ηfk(R)η(ρ)=(d2)+2ρ′′(ρ)(ρ),ζfk(R)ζ(ρ)=(d4)2ρ′′′(ρ)′′(ρ).\begin{split}\eta_{f_{k}(R)}\to\eta_{*}(\rho)&=-(d-2)+2\rho\frac{\mathcal{F}_{*}^{\prime\prime}(\rho)}{\mathcal{F}_{*}^{\prime}(\rho)}\,,\\ \zeta_{f_{k}(R)}\to\zeta_{*}(\rho)&=(d-4)-2\rho\frac{\mathcal{F}_{*}^{\prime\prime\prime}(\rho)}{\mathcal{F}_{*}^{\prime\prime}(\rho)}\,.\end{split} (77)

By substituting the fixed point values uu_{*}, given in Eq. (56), the results (73) and (74) allow to give the scaling dimension γn(u)m\gamma_{n}{}^{m}(u_{*}) at the fixed point.

Appendix B The flow equation for functional f(R)f(R)-truncations

For completeness, we reproduce the functional f(R)-equation obtained in Machado:2007ea .

384π2(tk+4k2ρk)=[5ρ2θ(1ρ3)(12+4ρ6190ρ2)][1ρ3]1+10ρ2θ(1ρ3)+[10ρ2θ(1ρ4)ρ2θ(1+ρ4)(36+6ρ6760ρ2)][1ρ4]1+[ηf(105ρ27136ρ2+72494536ρ3)+(6020ρ27118ρ2)][1+kkρ3]1+5ρ22[ηf((1+ρ3)θ(1+ρ3)+(2+ρ3)θ(1+ρ6))+2θ(1+ρ3)+4θ(1+ρ6)][1+kkρ3]1+[kηf(6+3ρ+2960ρ2+371512ρ3)+(tk′′2ρk′′′)(279120ρ22930ρ31813360ρ4)+k′′(216915ρ22915ρ3)+k(36+12ρ+2930ρ2)][2k+3k(123ρ)+9k′′(1ρ3)2]1,\begin{split}&384\pi^{2}\left(\partial_{t}{\mathcal{F}}_{k}+4{\mathcal{F}}_{k}-2\rho{\mathcal{F}}^{\prime}_{k}\right)=\vspace*{20cm}\\ &\quad\Big{[}5\rho^{2}\theta\left(1-\tfrac{\rho}{3}\right)-\left(12+4\,\rho-\tfrac{61}{90}\,\rho^{2}\right)\Big{]}\Big{[}1-\tfrac{\rho}{3}\Big{]}^{-1}+10\,\rho^{2}\,\theta\left(1-\tfrac{\rho}{3}\right)\\ &+\Big{[}10\,\rho^{2}\,\theta(1-\tfrac{\rho}{4})-\rho^{2}\,\theta(1+\tfrac{\rho}{4})-\left(36+6\,\rho-\tfrac{67}{60}\,\rho^{2}\right)\Big{]}\Big{[}1-\tfrac{\rho}{4}\Big{]}^{-1}\\ &+\Big{[}\eta_{f}\,\left(10-5\rho-\tfrac{271}{36}\rho^{2}+\tfrac{7249}{4536}\rho^{3}\right)+\left(60-20\rho-\tfrac{271}{18}\rho^{2}\right)\Big{]}\left[1+\tfrac{{\mathcal{F}}_{k}}{{\mathcal{F}}^{\prime}_{k}}-\tfrac{\rho}{3}\right]^{-1}\\ &+\frac{5\rho^{2}}{2}\,\Big{[}\eta_{f}\,\left((1+\tfrac{\rho}{3})\theta(1+\tfrac{\rho}{3})+(2+\tfrac{\rho}{3})\theta(1+\tfrac{\rho}{6})\right)+2\theta(1+\tfrac{\rho}{3})+4\theta(1+\tfrac{\rho}{6})\Big{]}\left[1+\tfrac{{\mathcal{F}}_{k}}{{\mathcal{F}}^{\prime}_{k}}-\tfrac{\rho}{3}\right]^{-1}\\ &+\Big{[}{\mathcal{F}}^{\prime}_{k}\,\eta_{f}\,\left(6+3\rho+\tfrac{29}{60}\rho^{2}+\tfrac{37}{1512}\,\rho^{3}\right)+\left(\partial_{t}{\mathcal{F}}^{\prime\prime}_{k}-2\rho{\mathcal{F}}^{\prime\prime\prime}_{k}\right)\left(27-\tfrac{91}{20}\rho^{2}-\tfrac{29}{30}\,\rho^{3}-\tfrac{181}{3360}\,\rho^{4}\right)\\ &+{\mathcal{F}}^{\prime\prime}_{k}\left(216-\tfrac{91}{5}\rho^{2}-\tfrac{29}{15}\rho^{3}\right)+{\mathcal{F}}^{\prime}_{k}\left(36+12\rho+\tfrac{29}{30}\rho^{2}\right)\Big{]}\Big{[}2{\mathcal{F}}_{k}+3{\mathcal{F}}^{\prime}_{k}(1-\tfrac{2}{3}\rho)+9{\mathcal{F}}^{\prime\prime}_{k}(1-\tfrac{\rho}{3})^{2}\Big{]}^{-1}\,,\end{split} (78)

where we used the short-hand notation

ηf=1k(tk+2k2ρk′′).\eta_{f}=\frac{1}{{\mathcal{F}}^{\prime}_{k}}\,\left(\partial_{t}{\mathcal{F}}^{\prime}_{k}+2{\mathcal{F}}^{\prime}_{k}-2\rho{\mathcal{F}}^{\prime\prime}_{k}\right)\,. (79)

This equation forms the basis for evaluating BFRGEB_{*}^{\rm FRGE} in Sec. 3.2.

References