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

Spinfoams, 𝜸\gamma-duality and parity violation in primordial gravitational waves

Eugenio Bianchi ebianchi@psu.edu    Monica Rincon-Ramirez mramirez@psu.edu Institute for Gravitation and the Cosmos, The Pennsylvania State University, University Park, Pennsylvania 16802, USA Department of Physics, The Pennsylvania State University, University Park, Pennsylvania 16802, USA
Abstract

The Barbero-Immirzi parameter γ\gamma appears as a coupling constant in the spinfoam dynamics of loop quantum gravity and can be understood as a measure of gravitational parity violation via a duality rotation. We investigate an effective field theory for gravity and a scalar field, with dynamics given by a γ\gamma-dual action obtained via a duality rotation of a parity-non-violating one. The resulting relation between the coupling constants of parity-even and parity-odd higher-curvature terms is determined by γ\gamma, opening the possibility of its measurement in the semiclassical regime. For a choice of γ\gamma-dual effective action, we study cosmic inflation and show that the observation of a primordial tensor polarization, together with the tensor tilt and the tensor-to-scalar ratio, provides a measurement of the Barbero-Immirzi parameter and, therefore, of the scale of discreteness of the quantum geometry of space.

I Introduction

The Barbero-Immirzi parameter γ\gamma plays a central role in Loop Quantum Gravity (LQG) [1, 2, 3]. The parameter sets the typical scale of the spectrum of geometric observables such as the volume of a quantum of space, the area of the interface between two quanta and the length of a curvature defect line [4, 5, 6, 7, 8]. The elementary area operator has eigenvalues

Aj=8πGγj(j+1),j{0,12, 1,32,}.A_{j}=8\pi G\hbar\,\gamma\,\sqrt{j(j+1)}\,,\qquad j\in\left\{0,\,\tfrac{1}{2},\,1,\,\tfrac{3}{2},\ldots\right\}\,. (1)

In particular, the theory predicts an area gap a=43πγP2a_{*}=4\sqrt{3}\,\pi\,\gamma\,\ell_{P}^{2}\,, where P=G/c31.6×1035m\ell_{P}=\sqrt{G\hbar/c^{3}}\simeq 1.6\times 10^{-35}\,\mathrm{m} is the Planck length and the Barbero-Immirzi parameter γ>0\gamma>0 is a dimensionless coupling constant that, in principle, is to be determined experimentally. At the classical level [9, 10, 11, 12], the parameter γ\gamma can be understood as a coupling constant in the Einstein-Cartan-Holst action for gravity. This action is invariant under orientation-preserving diffeomorphisms but not under orientation-reversing ones, and therefore is parity violating:

S=116πG12ϵIJKLeIeJFKL1γeIeJFIJ,S=\frac{1}{16\pi G}\!\int\frac{1}{2}\epsilon_{IJKL}\,e^{I}\wedge e^{J}\wedge F^{KL}-\frac{1}{\gamma}\,e_{I}\wedge e_{J}\wedge F^{IJ}\,, (2)

where the first term in the integral is the parity-odd Einstein-Cartan density and the second is the parity-even Holst density. At the quantum level, spinfoams provide a non-perturbative definition of the covariant dynamics of LQG [13]. The spinfoam dynamics is not defined in terms of an action for fields on a manifold, but instead by a spinfoam vertex amplitude. The Engle-Pereira-Rovelli-Livine (EPRL) [14] vertex amplitude WγW_{\gamma} depends on the Barbero-Immirzi parameter γ\gamma and belongs to a one-parameter family of spinfoam models related by a duality rotation with an angle θ\theta. In the limit θ0\theta\to 0 at fixed eigenvalues of the area, the EPRL model reduces to the Barrett-Crane model (BC) [15], which is parity invariant. On the other hand, the EPRL model with finite value of γ\gamma is parity-violating [16] and defined by

tanθ=1γ.\tan\theta=\frac{1}{\gamma}\,. (3)

At the semiclassical level, this one-parameter family of theories can be understood in terms of a duality rotation of the curvature

FIJ\displaystyle F^{IJ}\quad 𝜃+cosθF+IJsinθFIJ,\displaystyle\overset{\theta}{\longrightarrow}\quad+\cos\theta\;\,F{{}^{IJ}}\;+\;\sin\theta\;\;{}^{*}\!F^{IJ}\,, (4)
FIJ\displaystyle{}^{*}\!F^{IJ}\quad 𝜃sinθFIJ+cosθFIJ,\displaystyle\overset{\theta}{\longrightarrow}\quad-\sin\theta\;\>F^{IJ}\;+\;\cos\theta\;\;{}^{*}\!F^{IJ}\,,

where FIJ=12ϵIJKLFKL{}^{*}\!F_{IJ}=\frac{1}{2}\epsilon_{IJKL}\,F^{KL} is the Hodge dual in the internal Lorentz indices. Starting from an effective action for gravity that is parity non-violating, we can produce a parity-violating effective action via a duality rotation of the curvature. For instance, the action (2) can be obtained by a duality rotation of the Einstein-Cartan density with angle θ\theta satisfying the condition (3). We note that the duality rotation is not a symmetry of the action, or of its solutions [17, 18, 19], but a relation between theories with different values of their coupling constants.

When the effective action is organized as a derivative expansion [20], each term in the expansion is a Lagrangian density that comes with a coupling constant which, in principle, needs to be fixed by observations. Remarkably, the requirement that the effective action is obtained by a duality rotation of a parity non-violating effective action introduces a non-trivial relation between these coupling constants — a single parameter γ\gamma sets the scale of all parity violating terms in the action, modeling at the semiclassical level an exact property of the spinfoam dynamics WγW_{\gamma}. An effective field theory that satisfies this requirement is described by an action SγS_{\gamma} that we call γ\gamma-dual.

While a top-down derivation of the effective action directly from the non-perturbative spinfoam dynamics WγW_{\gamma} is still missing, the condition of γ\gamma-duality allows us to relate coupling constants of parity-even and parity-odd terms in the effective action SγS_{\gamma}. Therefore the observation of gravitational parity violation would provide an indirect measurement of the Barbero-Immirzi parameter γ\gamma, and therefore of the LQG area gap. In this paper we investigate a γ\gamma-dual model of cosmological inflation and show how the spectrum of primordial gravitational waves depends on the coupling constant γ\gamma.

The effect of gravitational parity violation in primordial inflation has been studied for the first time in [21] where a Chern-Simons modification of General Relativity was considered and a non-vanishing polarization Π(AT+AT)/(AT++AT)\Pi\equiv(A_{T+}-A_{T-})/(A_{T+}+A_{T-}) is shown to arise for the amplitude AT±A_{T\pm} of primordial tensor modes with circular polarization (±)(\pm). We refer to [22] for a review and to [23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38] for older and more recent developments. Our analysis is motivated by a relation between parity violation in primordial gravitational waves and the Barbero-Immirzi parameter in LQG originally proposed in [26]. We develop this proposal by introducing a new crucial ingredient — spinfoam γ\gamma-duality. In general, at the quadratic order in curvature, the Lagrangian for gravity and an inflaton scalar field ϕ\phi can include both a parity-even term LGBL_{GB} proportional to the Gauss-Bonnet density and a parity-odd term LCSL_{CS} proportional to the Pontryagin density (also know as Chern-Simons term) [27, 28], with

LGB=\displaystyle L_{GB}= fGB(ϕ)(RμνρσRμνρσ4RμνRμν+R2),\displaystyle\,f_{GB}(\phi)\,(R_{\mu\nu\rho\sigma}R^{\mu\nu\rho\sigma}\!-\!4R_{\mu\nu}R^{\mu\nu}\!+\!R^{2})\,, (5)
LCS=\displaystyle L_{CS}= fCS(ϕ)(1gϵμνρσRαRββμν)αρσ.\displaystyle\,f_{CS}(\phi)\;\big{(}\tfrac{-1}{\sqrt{-g}}\epsilon^{\mu\nu\rho\sigma}\,R^{\alpha}{}_{\beta\mu\nu}\,R^{\beta}{}_{\alpha\rho\sigma}\big{)}\,. (6)

The two functions fGB(ϕ)f_{GB}(\phi) and fCS(ϕ)f_{CS}(\phi) are in principle independent. The requirement of γ\gamma-duality of the action imposes a relation between the two functions: fGB(ϕ)/fCS(ϕ)=γ1γf_{GB}(\phi)/f_{CS}(\phi)=\gamma-\frac{1}{\gamma}. In this paper we show that, in a γ\gamma-dual model of slow-roll inflation, the asymmetry Π\Pi of the amplitude of circular polarizations, the tensor-to-scalar ratio r(AT++AT)/ASr\equiv(A_{T+}+A_{T-})/A_{S} and the tensor tilt nTn_{T} are related to the parameter γ\gamma by the equation

1γγ=π8r+8nTΠ.\frac{1}{\gamma}-\gamma=\frac{\pi}{8}\,\frac{r+8\,n_{T}}{\Pi}. (7)

Therefore the observation of a primordial parity violation Π0\Pi\neq 0 and a violation of the (r,nT)(r,n_{T}) ‘consistency relations’, r8nTr\neq-8\,n_{T}, would provide a measurement of the Barbero-Immirzi parameter γ\gamma, and thus fix the scale aa_{*} of the area gap.

The analysis presented here is complementary to the framework of loop quantum cosmology [39, 40, 41, 42, 43] where Planck scale effects on the background dynamics and on metric perturbations during the pre-inflationary era are found to leave phenomenological imprints in cosmic microwave background observables. The analysis is also complementary to the approach of spinfoam cosmology [44, 45, 46] where one investigates the cosmological regime of the full spinfoam dynamics. The phenomenological effects considered in this paper have their origin in the inflationary era, where the curvature is already far from the Planck scale and an effective field theory motivated by spinfoams can be used. In this sense, the analysis follows the same logic used in the study of parity violation in the coupling of fermions to gravity and torsion in an effective field theory [47, 48], with the additional new ingredient of γ\gamma-duality introduced here.

Specifically, we adopt the point of view that γ\gamma is a coupling constant that needs to be determined experimentally, as any other coupling constant. This viewpoint is adopted also in recent analysis in loop quantum cosmology [49, 50], where observational bounds on the area gap aa_{*} — and therefore on γ\gamma — are investigated. This point of view differs from the one (adopted for instance in [42]) where γ\gamma is set to a fixed exact value γ0=log(2)/π3\gamma_{0}=\log(2)/\pi\sqrt{3} [51] or γM\gamma_{M} [52], which is obtained by imposing that the microscopic entropy of the horizon-area ensemble matches the thermodynamics entropy of a black hole of the same horizon area. Other analysis of the derivation of the Bekenstein-Hawking entropy formula show that the coefficient of the area law is determined by semiclassical infrared phenomena and is independent of the value of the Barbero-Immirzi parameter [53, 54, 55, 56], and therefore needs to be determined experimentally.

The paper is organized as follows. In section II we describe γ\gamma-duality in spinfoams, both at the level of the non-perturbative definition of the vertex amplitude and at the semiclassical level. In section III we discuss the link between parity, orientation-reversing diffeomorphisms and internal Lorentz reversals, and describe how different gravitational terms in the effective action transform. In section IV we use duality rotations to construct a spinfoam-motivated effective action for inflation and discuss the assumptions of the model. In section V we compute the primordial power spectrum of primordial gravitational waves, and in section VI we discuss prospects for observations.

II 𝜸\gamma-Duality in Spinfoams

We briefly discuss the EPRL spinfoam model [1, 13], illustrating its parity-violating nature [16, 57, 58, 59] and the role of γ\gamma-duality both in its non-perturbative definition and in the semiclassical limit.

The Barbero-Immirzi parameter appears as a dimensionless coupling constant in the Einstein-Cartan-Holst action SECHS_{ECH}, (2). This action can be obtained starting from the Einstein-Cartan action SECS_{EC},

SEC=12κ12ϵIJKLeIeJFKL\displaystyle S_{EC}=\frac{1}{2\kappa}\,\int\frac{1}{2}\epsilon_{IJKL}\,e^{I}\wedge e^{J}\wedge F^{KL} (8)
𝜃SECH=cosθ2κ12ϵIJKLeIeJFKL\displaystyle\overset{\theta}{\longrightarrow}\quad S_{ECH}=\frac{\cos\theta}{2\kappa}\,\int\frac{1}{2}\epsilon_{IJKL}\,e^{I}\wedge e^{J}\wedge F^{KL}
tanθeIeJFIJ,\displaystyle\hskip 102.00012pt-\tan\theta\;e_{I}\wedge e_{J}\wedge F^{IJ}\,, (9)

via the duality rotation (4) of the curvature. Comparing the expressions (2) and (9), we observe that the Barbero-Immirzi parameter is related to the duality rotation angle θ\theta by the relation (3) and κ\kappa is related to the observed value of Newton’s constant GG by 8πG=κ/cosθ8\pi G=\kappa/\cos\theta.


In spinfoam quantum gravity one finds a similar relation at the non-perturbative level. The relation can be expressed in terms of the labels (p,j)(p,j) of the unitary irreducible representations of the principal series of the Lorentz group SL(2,)SL(2,\mathbb{C}),

gSL(2,)D(p,j)(g).g\in SL(2,\mathbb{C})\qquad\longrightarrow\quad D^{(p,j)}(g)\,. (10)

The Engle-Pereira-Rovelli-Livine spinfoam model (EPRL) with Barbero-Immirzi parameter γ\gamma [14] is defined by the representation D(γj,j)(g)D^{(\gamma j,j)}(g), while the Barrett-Crane model (BC) [15] by the representation D(p0,0)(g)D^{(p_{0},0)}(g). The two models are related by the duality rotation

p=cosθp0,j=sinθp0,p\;=\;\cos\theta\;p_{0}\;,\qquad j\;=\;\sin\theta\;p_{0}\;, (11)

where the angle θ\theta satisfies (3), i.e.,

p=γj.p\,=\,\gamma\,j\,. (12)

The BC model is obtained from the EPRL model as the limit θ0\theta\to 0 at fixed p0p_{0}. Equivalently, this duality rotation defines a one-parameter family of parity-violating EPRL models starting from the parity-non-violating BC model as a seed. Specifically, the spinfoam wedge amplitude 𝒜w\mathcal{A}_{w} for a coherent boundary state labeled by the spinors ζ\zeta and ζ\zeta^{\prime} takes the form [60]

𝒜w(g,ζ,ζ′′)=[jw,ζ|D(pw,jw)(g)|jw,ζ′′=P1𝑑μeiSw,\mathcal{A}_{w}(g,\zeta^{\prime},\zeta^{\prime\prime})=[\,j_{w},\zeta^{\prime}|D^{(p_{w},j_{w})}(g)|j_{w},\zeta^{\prime\prime}\rangle=\!\!\int_{\mathbb{C}P^{1}}\!\!\!d\mu\;\;\mathrm{e}^{\,\mathrm{i}S_{w}}, (13)

where the wedge action SwS_{w} consists of the sum of two terms proportional to the quantum numbers pwp_{w} and jwj_{w} that label the representation, i.e., Sw=pwΘw+jwχwS_{w}=p_{w}\,\Theta_{w}\;+\;j_{w}\,\chi_{w}. In the semiclassical limit for a Lorentzian 44-simplex, a saddle-point analysis shows that, for p0p_{0}\to\infty, the integral is dominated by the exponential of the sum of the wedge actions evaluated at each saddle point,

Sv=wS¯w=wpwΘ¯w+wjwχ¯w.S_{v}=\sum_{w}\bar{S}_{w}\;=\;\sum_{w}p_{w}\,\bar{\Theta}_{w}\;+\;\sum_{w}j_{w}\,\bar{\chi}_{w}\,. (14)

The two terms appearing in the semiclassical action for a spinfoam vertex can be understood as a discretization on a Lorentzian 44-simplex of the two terms in the Einstein-Cartan-Holst action (9). The first of the two terms reproduces the Regge action for a Lorentzian 44-simplex [61],

18πGSRegge=wAw8πGΘ¯w,\frac{1}{8\pi G\hbar}S_{\mathrm{Regge}}\;=\;\sum_{w}\frac{A_{w}}{8\pi G\hbar}\bar{\Theta}_{w}\,, (15)

where AwA_{w} is the area of the triangle where the wedge ww hinges, and Θ¯w\bar{\Theta}_{w} is the boost angle between the two space-like tetrahedra that define the wedge. This relation is consistent with the area spectrum in loop quantum gravity in the large spin limit,

Aw=8πGpw= 8πGγjw.A_{w}=8\pi G\hbar\;p_{w}\;=\;8\pi G\hbar\,\gamma\,j_{w}\,. (16)

On the other hand, the second term in (14) includes a twist angle χ¯w\bar{\chi}_{w} that depends on the phase of the coherent boundary state spinors and does not affect the dynamics of the model. This term is analogous to the Holst term in (9) and has the same transformation properties under parity111Their parity can be read for instance in the asymptotics of the vertex amplitude for Lorentzian boundary data, where a sum of saddle points results in the two terms of different parity Ave+iwpwΘ¯w+iwjwχ¯w+eiwpwΘ¯w+iwjwχ¯wA_{v}\sim\mathrm{e}^{+\mathrm{i}\sum_{w}p_{w}\,\bar{\Theta}_{w}\;+\;\mathrm{i}\sum_{w}j_{w}\,\bar{\chi}_{w}}+\mathrm{e}^{-\mathrm{i}\sum_{w}p_{w}\,\bar{\Theta}_{w}\;+\;\mathrm{i}\sum_{w}j_{w}\,\bar{\chi}_{w}}.. Note also that the Barrett-Crane model is obtained as the limit θ0\theta\to 0 where this second term is not present. When changing the angle θ\theta, the geometric relation between the area AwA_{w} and the representation pwp_{w} can be preserved by rescaling κ\kappa by cosθ\cos\theta as discussed after (9) at the level of the classical action.

The argument described above shows that the duality rotation (4) relates the semiclassical limit of spinfoam models with different values of the Barbero-Immirzi parameter. Remarkably, this is an exact relation that holds beyond the semiclassical limit p01p_{0}\gg 1. The defining equations of the EPRL spinfoam model with Barbero-Immirzi parameter γ\gamma is the linear simplicity constraint [1]

K=γL\vec{K}=\gamma\vec{L} (17)

where K\vec{K} and L\vec{L} are the generators of boosts and rotations for the Lorentz group.222Specifically, let us denote JIJJ^{IJ} the generators of the Lorentz group SL(2,)SL(2,\mathbb{C}) and introducing an orthonormal frame eμI=(uI,ξiI)e^{I}_{\mu}=(u^{I},\xi_{i}^{I}) with ηIJuIuJ=1\eta_{IJ}u^{I}u^{J}=-1. We can then define the generators of the SU(2)SU(2) rotations group Li=12ϵIJKLJIJuKξiLL_{i}=\frac{1}{2}\epsilon_{IJKL}J^{IJ}u^{K}\xi^{L}_{i} that preserve the timelike direction uIu^{I}, and the generators of boosts Ki=JIJξiIuJK_{i}=J_{IJ}\xi^{I}_{i}u^{J}. In terms of the self-dual and the anti-self-dual generators,

Ji(±)=Li±iKi2,J_{i}^{(\pm)}=\frac{L_{i}\pm\mathrm{i}\,K_{i}}{2}\,, (18)

the linear simplicity constraint takes the form

Ji(±)=1±iγ1+γ2Ji(0)J_{i}^{(\pm)}=\frac{1\pm\mathrm{i}\,\gamma}{\sqrt{1+\gamma^{2}}}\,J^{(0)}_{i}\, (19)

where Ji(0)J^{(0)}_{i} is defined by the parity-even limit γ\gamma\to\infty. We can now introduce a duality rotation [62]

Ji(±)𝜃e±iθJi(±),J_{i}^{(\pm)}\;\overset{\theta}{\longrightarrow}\;\;\mathrm{e}^{\pm\mathrm{i}\theta}\,J_{i}^{(\pm)}\,, (20)

and notice again that the EPRL and the BC models are consistently related by a duality rotation with angle θ\theta given by (3). In a (formal) connection representation Ψ[ωi(+),ωi()]\Psi[\omega_{i}^{(+)},\omega_{i}^{(-)}] where the self-dual/anti-selfdual generators act as a derivative, Ji(±)Ψ[ωi(+),ωi()]=iδδωi±Ψ[ωi(+),ωi()]J_{i}^{(\pm)}\Psi[\omega_{i}^{(+)},\omega_{i}^{(-)}]=-\mathrm{i}\frac{\delta}{\delta\omega_{i}^{\pm}}\Psi[\omega_{i}^{(+)},\omega_{i}^{(-)}], the duality transformation relates the amplitude of states for theories with different value of the Barbero-Immirzi parameter γ\gamma, with the requirement that, in the limit γ\gamma\to\infty at fixed elementary areas, the theory is parity invariant. Any spinfoam vertex that satisfies this condition, and therefore implements the relation (9) at the non-perturbative level, is called γ\gamma-dual. The EPRL model is a specific and concrete proposal in this class. In the rest of this paper we take this property, γ\gamma-duality, as fundamental and we explore its consequences in a semiclassical limit described by an effective action.

III Parity and Orientation-Reversing Diffeomorphisms

In this section we discuss the notion of parity in a general relativistic theory where geometry is dynamical. In particular we show how orientation-reversing diffeomorphisms generalize the notion of parity defined in a special relativistic theory, and clarify the link between orientation-reversing diffeomorphisms and orientation reversals in the internal Lorentz group in the Einstein-Cartan formalism. We then describe how different gravitational terms in the effective action transform.

In 3-dimensional Euclidean space, a parity transformation P:{x,y,z}{x,y,z}P\,:\,\{x,y,z\}\mapsto\{-x,-y,-z\} is defined as the simultaneous reversal of the three spatial Cartesian axes, that is the product of the three spatial-inversion operators with, e.g., Px:xxP_{x}:\,x\mapsto-x. Similarly, a time reversal T:ttT:\,t\mapsto-t is a flip of the temporal axis. In a special relativistic theory, these specific transformations clearly depend on a choice of an inertial rest frame in Minkowski space. It is then useful to think of them in terms of Lorentz transformations that are not connected to the identity. The Lorentz group O(1,3)O(1,3) is the group of transformations ΛIJ\Lambda^{I}{}_{J} that preserve the Minkowski metric ηIJ\eta_{IJ}. The parameter space of Lorentz transformations splits into four connected components O(1,3)L+L+LLO(1,3)\simeq L_{+}^{\uparrow}\cup L_{+}^{\downarrow}\cup L_{-}^{\uparrow}\cup L_{-}^{\downarrow}, with the subgroup of proper orthochronous Lorentz transformations O0(1,3)=SO(1,3)L+O_{0}(1,3)=SO^{\uparrow}(1,3)\simeq L_{+}^{\uparrow} being the components connected to the identity [63, 64]. We can then consider Lorentz transformations modulo transformations connected to the identity,

O(1,3)/O0(1,3)={𝟙,P,T,PT},O(1,3)/O_{0}(1,3)\;=\;\{\mathds{1},\,P,\,T,\,PT\}\,, (21)

with the quotient space generated by parity PP and time reversal TT. Orientation reversals of Minkowski space are Lorentz transformations Λ\Lambda with parameters in L+L=TL+PL+L_{+}^{\downarrow}\cup L_{-}^{\uparrow}\,=\,TL_{+}^{\uparrow}\cup PL_{+}^{\uparrow} and are defined by det(Λ)=1\det(\Lambda)=-1.

The notions of parity and time reversal described above require a fixed background metric such as the Minkowski metric. Here we are interested in gravitational parity violation, where the metric and its causal structure are not fixed as a background structure but are dynamical instead. Before defining orientation reversals for a spacetime manifold, let us start from a vector space, a linear space of dimension dd that is not equipped with any additional structure. In this case, given an ordered basis of dd linearly independent vectors EI={E1,,Ed}E_{I}=\{E_{1},\ldots,E_{d}\}, we can consider a change of basis given by a matrix MM. The requirement that E~I=MIEJJ\tilde{E}_{I}\,=\,M_{I}{}^{J}\,E_{J} is still a linear basis imposes that detM0\det M\neq 0. Therefore we have two distinct equivalence classes of bases: If det(M)>0\det(M)>0 then the two bases belong to the same class; if det(M)<0\det(M)<0, they belong to different classes. Each one of such equivalence classes is an orientation Ξ\Xi for the vector space.

In the case of a differentiable dd-dimensional manifold \mathcal{M}, an orientation Ξ\Xi can be equivalently defined as either [65, 66, 67]: (i) A continuous point-wise orientation on the manifold; (ii) An equivalence class of nowhere-vanishing top-forms (or dd-forms in a dd-dimensional manifold); or, (iii) An equivalence class of oriented atlases. A manifold is called orientable if it is possible to make a choice of orientation. If \mathcal{M} is connected and orientable, there are again only two possible orientations Ξ=±1\Xi=\pm 1.

Definition (i) is the generalization of the concept of orientation of a vector space introduced above, where one considers a frame field EIE_{I} (i.e., a set of bases EI(p)E_{I}(p), one at each point pp\in\mathcal{M}), with the additional condition that the frame must have a point-wise continuous class: at every point pp there is a neighborhood with the same choice of orientation Ξp\Xi_{p} of its tangent space.333We note that this definition of orientation of a manifold does not require the existence of a global continuous basis. The frame can be discontinuous, as long as the point-wise orientation is continuous.

Definition (ii) is the one that is of the most relevance here. Let us discuss briefly its equivalence to (i). Considering a nowhere-vanishing top-form Ω\Omega, and an orientation Ξ=[EI]\Xi=[E_{I}] for which EIE_{I} is a representing frame, we can check if the inner product Ω,E1Ed\langle\Omega,E_{1}\cdots E_{d}\rangle is either positive or negative as, by definition, it cannot be zero. This defines an equivalence relation for nowhere-vanishing top-forms, splitting them into two disjoint sets. The orientation Ξ\Xi of the manifold \mathcal{M} is given by an equivalence class of smooth nowhere-vanishing top-forms Ω\Omega and we have Ξ=[EI]=[Ω]\Xi=[E_{I}]=[\Omega] with Ω,E1Ed>0\langle\Omega,E_{1}\cdots E_{d}\rangle>0.

Definition (iii) in terms of an oriented atlas is specifically useful once we introduce coordinates xμx^{\mu} and dynamical fields over the manifold, as done in General Relativity. We set d=4d=4 and assume that the 44-dimensional manifold \mathcal{M} representing spacetime is orientable. The orientation Ξ=[μ]\Xi=[\partial_{\mu}] in each chart of the atlas is given by the coordinate frame μ=/xμ\partial_{\mu}=\partial/\partial x^{\mu}. A spacetime diffeomorphism φDiff()\varphi\in\mathrm{Diff}(\mathcal{M}) is then a smooth function yμ=φμ(x)y^{\mu}=\varphi^{\mu}(x) with smooth inverse. Under a spacetime diffeomorphism, the nowhere vanishing 44-form dx0dx3dx^{0}\wedge\cdots\wedge dx^{3} transforms as

dy0dy1dy2dy3=det(Jμν)dx0dx1dx2dx3,dy^{0}\wedge dy^{1}\wedge dy^{2}\wedge dy^{3}\,=\,\det(J^{\mu}{}_{\nu})\,dx^{0}\wedge dx^{1}\wedge dx^{2}\wedge dx^{3}\,, (22)

where Jμ=νyμ/xνJ^{\mu}{}_{\nu}=\partial y^{\mu}/\partial x^{\nu} is the Jacobian of the transformation. An orientation-preserving diffeomorphism is characterized by det(J)>0\det(J)>0, an orientation-reversing diffeomorphism by det(J)<0\det(J)<0. Diffeomorphisms that are connected to the identity, denoted by Diff0()\mathrm{Diff}_{0}(\mathcal{M}), form a subgroup and the action of infinitesimal diffeomorphisms on fields is represented by the Lie derivative ξ\mathcal{L}_{\xi} with respect to the vector field ξμ\xi^{\mu}. If we assume that the spacetime manifold \mathcal{M} is orientable and has a topology that is trivial [68], we have then that the mapping class group of spacetime diffeomorphisms modulo transformations connected to the identity,

Diff()/Diff0()={𝟙,},\mathrm{Diff}(\mathcal{M})/\mathrm{Diff}_{0}(\mathcal{M})\;=\;\{\mathds{1},\mathcal{R}\}\,, (23)

is generated by the manifold orientation reversal :ΞΞ\mathcal{R}:\Xi\mapsto-\Xi. As we have introduced no Lorentzian metric up to this point, there is no distinction between spatial parity and time reversal, only a notion of orientation reversal \mathcal{R}.

In the Einstein-Cartan formulation of General Relativity, the fundamental dynamical variables are a Lorentz connection ωμIJ(x)dxμ\omega_{\mu}^{IJ}(x)dx^{\mu} and a coframe field eμI(x)dxμe^{I}_{\mu}(x)dx^{\mu}. The Lorentz connection allows us to parallel transport Lorentz fields. The coframe field allows us to define a Lorentzian metric as a derived quantity. The ordered basis EI=EIμ(x)μE_{I}=E_{I}^{\mu}(x)\partial_{\mu} discussed earlier, the frame field, is also a derived quantity defined by eI,EJ=δIJ\langle e^{I},E_{J}\rangle=\delta^{I}{}_{J}, i.e., the change of basis EIμE_{I}^{\mu} from the coordinate frame μ\partial_{\mu} to the non-coordinate frame EIE_{I} is given by the inverse of the coframe field, EIμ=(eμI)1E_{I}^{\mu}=(e_{\mu}^{I})^{-1}. The requirement that the two basis represent the same orientation implies that the determinant of the coframe field is positive,

Ξ=[μ]=[EI]det(eμI)>0.\Xi=[\partial_{\mu}]=[E_{I}]\quad\Longrightarrow\quad\det(e_{\mu}^{I})>0\,. (24)

Equivalently, we can consider the volume 44-form

Ωv=14!ϵIJKLeIeJeKeL,\Omega_{\mathrm{v}}\,=\,\frac{1}{4!}\,\epsilon_{IJKL}\,e^{I}\wedge e^{J}\wedge e^{K}\wedge e^{L}\,, (25)

together with the coordinate 44-form

d4x=14!ϵμνρσdxμdxνdxρdxσ,\differential^{4}x\,=\,\frac{1}{4!}\,\epsilon_{\mu\nu\rho\sigma}\,dx^{\mu}\wedge dx^{\nu}\wedge dx^{\rho}\wedge dx^{\sigma}\,, (26)

where ϵIJKL\epsilon_{IJKL} and ϵμνρσ\epsilon_{\mu\nu\rho\sigma} are alternating symbols (with ϵ0123=+1\epsilon_{0123}=+1). As they are both nowhere-vanishing top forms and Ωv=det(eμI)d4x\Omega_{\mathrm{v}}=\det(e_{\mu}^{I})\,\differential^{4}x, we have again

Ξ=[d4x]=[Ωv]det(eμI)>0.\Xi=[\differential^{4}x]=[\Omega_{\mathrm{v}}]\quad\Longrightarrow\quad\det(e_{\mu}^{I})>0\,. (27)

We turn now to the discussion of diffeomorphisms, and in particular of orientation-reversing diffeomorphisms, in the Einstein-Cartan formulation. Let us consider a Lorentz vector-valued one-form αI=αμI(x)dxμ\alpha^{I}=\alpha^{I}_{\mu}(x)dx^{\mu}. Given a spacetime diffeomorphism φDiff()\varphi\in\mathrm{Diff}(\mathcal{M}), we need to define a Kosmann lift to the principle Lorentz bundle [69, 70, 71, 72],

φ^O(1,3)Diff(),\hat{\varphi}\;\in\;O(1,3)_{\mathcal{M}}\rtimes\mathrm{Diff}(\mathcal{M})\,, (28)

that allows us to compare Lorentz vectors at different spacetime points, i.e.,

αμI(x)(Λφ(x))IανKK(φ(x))(Jφ(x))ν.μ\alpha^{I}_{\mu}(x)\quad\longrightarrow\quad(\Lambda_{\varphi}(x))^{I}{}_{K}\;\alpha^{K}_{\nu}(\varphi(x))\;(J_{\varphi}(x))^{\nu}{}_{\mu}\,. (29)

At the infinitesimal level, the lift of a spacetime diffeomorphism to Lorentz vectors is given by the Lorentz-covariant Lie derivative ^ξ\hat{\mathcal{L}}_{\xi}. This notion requires the use of the Lorentz connection ωμIJ\omega_{\mu}^{IJ}, with

^ξαμI=ξνναμI+ανIμξν+ξνωνJIαμJ.\hat{\mathcal{L}}_{\xi}\alpha^{I}_{\mu}\;=\;\xi^{\nu}\partial^{\vphantom{I}}_{\nu}\alpha^{I}_{\mu}+\alpha^{I}_{\nu}\partial^{\vphantom{I}}_{\mu}\xi^{\nu}\;+\;\xi^{\nu}\omega^{\,I}_{\nu\,J}\,\alpha^{J}_{\mu}\,. (30)

For large diffeomorphims that are not connected to the identity, the natural lift φ^\hat{\varphi} extends consistently to reversals \mathcal{R} of the orientation defined in a coordinate basis Ξ=[μ]=[d4x]\Xi=[\partial_{\mu}]=[\differential^{4}x] or equivalently in a non-coordinate basis Ξ=[EI]=[Ωv]\Xi=[E_{I}]=[\Omega_{\mathrm{v}}], i.e.,

^:\displaystyle\hat{\mathcal{R}}:\;\; [d4x][d4x],\displaystyle[\differential^{4}x]\,\mapsto-[\differential^{4}x]\,, (31)
[Ωv][Ωv].\displaystyle[\Omega_{\mathrm{v}}]\;\>\mapsto-[\Omega_{\mathrm{v}}]\,. (32)

This natural lift of the spacetime orientation reversal requires that the local Lorentz transformation Λφ\Lambda_{\varphi} has the same determinant sign as the Jacobian (Jφ)μ=νφμ/xν(J_{\varphi})^{\mu}{}_{\nu}=\partial\varphi^{\mu}/\partial x^{\nu},

det(Λφ)=sgndet(Jφ).\det(\Lambda_{\varphi})\;=\;\operatorname{sgn}\det(J_{\varphi})\,. (33)

Specifically, an orientation reversing diffeomorphism det(Jφ)<0\det(J_{\varphi})<0 always comes together with an internal space Lorentz reversal, det(Λφ)=1\det(\Lambda_{\varphi})=-1. In particular, we have that under a spacetime orientation-reversing diffeomorphism φ^\hat{\varphi}, the coframe field transforms as eμI(Λφ)IeνKK(Jφ)νμe_{\mu}^{I}\to(\Lambda_{\varphi})^{I}{}_{K}\;e^{K}_{\nu}\;(J_{\varphi})^{\nu}{}_{\mu} and, as a result, det(e)>0\det(e)>0 is sent to det(ΛφeJφ)>0\det(\Lambda_{\varphi}\,e\,J_{\varphi})>0 . With these preliminaries, we are ready to discuss how this lift provides a consistent definition of spacetime orientation reversals both in the first-order and in the second-order formulations of General Relativity.

Einstein’s general covariance is realized in General Relativity as the invariance of the action S=ΩS=\int\Omega under spacetime diffeomorphisms. While diffeomorphisms connected to the identity are assumed to be an exact symmetry also at the quantum level, transformation in the mapping class group (23) are expected to be only accidental approximate symmetries in the semiclassical regime. This is what happens for instance in 2+12+1 quantum gravity [73] and in LQG in four dimensions [16]. As we assume here that the effective action is invariant only under Diff0()\mathrm{Diff}_{0}(\mathcal{M}), it is useful to list the transformation properties of various 44-forms Ω\Omega that can appear in the action for gravity. We start from the coframe field eIe^{I}, the coframe torsion τI\tau^{I}, and the Lorentz curvature FIJF^{IJ},

τI=\displaystyle\tau^{I}= eI=deI+ωIeJJ,\displaystyle\;\,\nabla e^{I}\,=\,de^{I}+\omega^{I}{}_{J}\,e^{J}\,, (34)
FI=J\displaystyle F^{I}{}_{J}= dωI+JωIKωK,J\displaystyle\;\,d\omega^{I}{}_{J}+\omega^{I}{}_{K}\wedge\omega^{K}{}_{J}\,, (35)

and we raise and lower Lorentz indices I,J,I,J,\ldots with the internal Minkowski metric ηIJ\eta_{IJ}. The only 44-form that is Lorentz invariant, polynomial in eIe^{I} and ωIJ\omega^{IJ}, and contains no exterior derivatives is the volume 44-form Ωv\Omega_{\mathrm{v}} (25). With one single exterior derivative we have the Einstein-Cartan and the Holst 44-forms

ΩEC=\displaystyle\Omega_{EC}\;=\; 12ϵIJKLeIeJFKL,\displaystyle\;\frac{1}{2}\epsilon_{IJKL}\,e^{I}\wedge e^{J}\wedge F^{KL}\,, (36)
ΩH=\displaystyle\Omega_{H}\;\;=\; eIeJFIJ,\displaystyle\;e_{I}\wedge e_{J}\wedge F^{IJ}\,, (37)

With two exterior derivatives, we have the torsion-squared 44-form ΩTT\Omega_{TT}, the Gauss-Bonnet (also known as Euler) 44-form ΩGB\Omega_{GB} and the Chern-Simons (also known as Pontryagin) 44-form ΩCS\Omega_{CS} [74, 75, 76, 77, 78]:

ΩTT=\displaystyle\Omega_{TT}\;=\; eIeI,\displaystyle\;\nabla e_{I}\wedge\nabla e^{I}\,, (38)
ΩGB=\displaystyle\Omega_{GB}\;=\; 12ϵIJKLFIJFKL,\displaystyle\;\frac{1}{2}\epsilon_{IJKL}\,F^{IJ}\wedge F^{KL}\,, (39)
ΩCS=\displaystyle\Omega_{CS}\;=\; FIJFIJ.\displaystyle\;F_{IJ}\wedge F^{IJ}\,. (40)

Under orientation-reversing diffeomorphisms \mathcal{R} we have the following transformation properties:

Ωv\displaystyle\Omega_{\mathrm{v}}\;\;\;\quad Ωv,\displaystyle\overset{\mathcal{R}}{\longrightarrow}\quad-\Omega_{\mathrm{v}}\,, (41)
ΩEC\displaystyle\Omega_{EC}\quad ΩEC,\displaystyle\overset{\mathcal{R}}{\longrightarrow}\quad-\Omega_{EC}\,, (42)
ΩH\displaystyle\Omega_{H}\;\;\quad +ΩH,\displaystyle\overset{\mathcal{R}}{\longrightarrow}\quad+\Omega_{H}\,, (43)
ΩTT\displaystyle\Omega_{TT}\quad +ΩTT,\displaystyle\overset{\mathcal{R}}{\longrightarrow}\quad+\Omega_{TT}\,, (44)
ΩGB\displaystyle\Omega_{GB}\quad ΩGB,\displaystyle\overset{\mathcal{R}}{\longrightarrow}\quad-\Omega_{GB}\,, (45)
ΩCS\displaystyle\Omega_{CS}\quad +ΩCS.\displaystyle\overset{\mathcal{R}}{\longrightarrow}\quad+\Omega_{CS}\,. (46)

Therefore, the Einstein-Cartan action with a cosmological constant, SECΛ=12κ(ΩEC2ΛΩv)S_{EC\Lambda}=\frac{1}{2\kappa}\int(\Omega_{EC}-2\Lambda\,\Omega_{\mathrm{v}}) is odd under orientation reversals, SECΛSECΛS_{EC\Lambda}\to-S_{EC\Lambda}. On the other hand, including also a Holst term or a Chern-Simons term results in an action that is neither odd nor even under orientation reversals. We will call an action that is neither odd nor even under orientation-reversing diffeomorphisms a parity-violating action for short.


The conditions for a parity-violating action described above apply also to the metric formulation of General Relativity. Let us assume that the spacetime torsion vanishes, τI=eI=0\tau^{I}=\nabla e^{I}=0, and that there is no coupling to spinor fields that can source torsion. We can then formulate the theory purely in terms the spacetime metric gμν(x)g_{\mu\nu}(x) taken as a fundamental field, instead of considering it a derived quantity

gμν(x)=ηIJeμI(x)eνJ(x),g_{\mu\nu}(x)\,=\,\eta_{IJ}^{\vphantom{I}}\,e^{I}_{\mu}(x)e^{J}_{\nu}(x)\,, (47)

as done in the Einstein-Cartan formulation. Moreover the condition of zero torsion τI=0\tau^{I}=0 allows us to solve for the metric-compatible Lorentz connection that defines the Christoffel connection Γμνρ=Γμνρ(g)\Gamma^{\rho}_{\mu\nu}=\Gamma^{\rho}_{\mu\nu}(g) and the Riemann tensor Rμ(g)νρσR^{\mu}{}_{\nu\rho\sigma}(g). In the metric formulation, the volume 44-form (25) can be expressed as

Ωv=gd4x\Omega_{\mathrm{v}}=\sqrt{-g}\,\differential^{4}x (48)

in terms of the coordinate 44-form d4x=dx0dx1dx2dx3\differential^{4}x=dx^{0}\wedge dx^{1}\wedge dx^{2}\wedge dx^{3} and the volume density g=detgμν\sqrt{-g}=\sqrt{-\det g_{\mu\nu}}. Note that we have already assumed that the coordinate basis and the coframe field define the same orientation, and therefore g=det(eμI)>0\sqrt{-g}=\det(e_{\mu}^{I})>0. We can now write the 44-forms described earlier in terms of the metric, under the assumption of vanishing torsion. In particular we have that the Einstein-Cartan 44-form reduces to the Ricci scalar times the volume 44-form (also known as the Einstein-Hilbert density), while the Holst 44-form vanishes:

ΩEC(τ=0)=\displaystyle\Omega_{EC}^{(\tau=0)}\;=\; Rgd4x,\displaystyle\;R\,\sqrt{-g}\,\differential^{4}x\,, (49)
ΩH(τ=0)=\displaystyle\Omega_{H}^{(\tau=0)}\;=\;  0.\displaystyle\;0\,. (50)

The second equation follows from the identity ΩH=ΩTTΩNY\Omega_{H}=\Omega_{TT}-\Omega_{NY} where ΩNY=d(eIτI)\Omega_{NY}=d(e_{I}\wedge\tau^{I}) is the Nieh-Yan 44-form. Similarly, the Gauss-Bonnet and the Chern-Simons 44-forms in the absence of torsion reduce to

ΩGB(τ=0)=\displaystyle\Omega_{GB}^{(\tau=0)}= 12(RμνρσRμνρσ4RμνRμν+R2)gd4x,\displaystyle\;\tfrac{1}{2}\big{(}R_{\mu\nu\rho\sigma}R^{\mu\nu\rho\sigma}-4R_{\mu\nu}R^{\mu\nu}+R^{2}\big{)}\,\sqrt{-g}\,\differential^{4}x\,,
ΩCS(τ=0)=\displaystyle\Omega_{CS}^{(\tau=0)}= 14ϵμνρσRαRββμνd4αρσx.\displaystyle\;\tfrac{1}{4}\,\epsilon^{\mu\nu\rho\sigma}\,R^{\alpha}{}_{\beta\mu\nu}\,R^{\beta}{}_{\alpha\rho\sigma}\;\differential^{4}x\,. (51)

In the metric formulation, the Lagrangian LL associated to a 44-form Ω\Omega is defined as

Ω=Lgd4x.\Omega\,=\,L\,\sqrt{-g}\,\differential^{4}x\,. (52)

Orientation-reversing diffeomorphisms in the metric formulation do not require a lift from spacetime to the Lorentz bundle (28). The transformation properties of the metric 44-forms (49-51) coincide with (41-46). Note that, as d4xd4x\differential^{4}x\to-\differential^{4}x under orientation reversals, we have that the action S=ΩS=\int\Omega and the associated Lagrangian LL have opposite transformation properties

SSL±L,S\,\overset{\mathcal{R}}{\longrightarrow}\,\mp S\quad\Longrightarrow\quad L\,\overset{\mathcal{R}}{\longrightarrow}\,\pm L\,, (53)

i.e., the Lagrangian LL transforms as a scalar or a pseudoscalar. In particular, the Einstein-Hilbert action SEH=12κRgd4xS_{EH}=\frac{1}{2\kappa}\int R\,\sqrt{-g}\,\differential^{4}x is odd under orientation reversals, while the Einstein-Hilbert Lagrangian LEH=12κRL_{EH}=\frac{1}{2\kappa}R is even and therefore a scalar.

In this analysis, we have focused on orientation-reversing spacetime diffeomorphisms and their lift to the principle Lorentz bundle. We have not discussed purely internal Lorentz transformations that are Lorentz reversals, such as the internal parity PP and internal time reversal TT considered in [16]. These transformations change the determinant of the frame field eμIe_{\mu}^{I} without changing the orientation of the spacetime manifold, and therefore they do not have a corresponding transformation in the metric formulation.

In formal manipulations in metric variables, it is useful to adopt the metric-dependent antisymmetric tensor [79],

ε(g)μνρσ=1gϵμνρσ,\varepsilon_{(g)}^{\mu\nu\rho\sigma}=\frac{-1}{\sqrt{-g}}\,\epsilon^{\mu\nu\rho\sigma}\,, (54)

which transforms as a tensor and is defined in terms of the Levi-Civita alternating symbol ϵμνρσ\epsilon^{\mu\nu\rho\sigma} (with ϵ0123=+1\epsilon^{0123}=+1) appearing for instance in (51).

We include here a remark about the Gauss-Bonnet and Chern-Simons densities that will be relevant in the next section. The integral of ΩGB\Omega_{GB} and of ΩCS\Omega_{CS} defines topological invariants of a manifold as they are given by the exterior derivatives of 33-forms,

ΩGB=\displaystyle\Omega_{GB}= d(12ϵIJKLωIJ(FKL13ωKMωML)),\displaystyle\;d\big{(}\tfrac{1}{2}\epsilon_{IJKL}\,\omega^{IJ}\wedge\big{(}F^{KL}-\tfrac{1}{3}\omega^{K}{}_{M}\wedge\omega^{ML}\big{)}\big{)}, (55)
ΩCS=\displaystyle\Omega_{CS}= d(ωIJ(FIJ13ωIMωMJ)).\displaystyle\;d\big{(}\omega_{IJ}\wedge\big{(}F^{IJ}-\tfrac{1}{3}\omega^{I}{}_{M}\wedge\omega^{MJ}\big{)}\big{)}\,. (56)

In metric variables

LGB=μKGBμ\displaystyle L_{GB}=\nabla_{\mu}K_{GB}^{\mu} (57)
LCS=μKCSμ\displaystyle L_{CS}=\nabla_{\mu}K_{CS}^{\mu} (58)

where KGBμK_{GB}^{\mu} is given by [80]

KGBμ=ε(g)μνρσε(g)αβΓανγγδ(12Rδ+βρσ13ΓλρδΓβσλ),K_{GB}^{\mu}=\varepsilon_{(g)}^{\mu\nu\rho\sigma}\varepsilon_{(g)}^{\alpha\beta}{}_{\gamma\delta}\,\Gamma^{\gamma}_{\alpha\nu}\big{(}\tfrac{1}{2}R^{\delta}{}_{\beta\rho\sigma}+\tfrac{1}{3}\Gamma^{\delta}_{\lambda\rho}\Gamma^{\lambda}_{\beta\sigma}\big{)}\,, (59)

and KCSμK_{CS}^{\mu} is the Chern-Simons current, which in the absence of torsion can be written as [22, 23]

KCSμ=ε(g)μνρσΓνβα(ρΓσαβ+23ΓρλβΓσαλ),K_{CS}^{\mu}=\varepsilon_{(g)}^{\mu\nu\rho\sigma}\,\Gamma^{\alpha}_{\nu\beta}\big{(}\partial_{\rho}\Gamma^{\beta}_{\sigma\alpha}+\tfrac{2}{3}\Gamma^{\beta}_{\rho\lambda}\Gamma^{\lambda}_{\sigma\alpha}\big{)}, (60)

Even though by themselves they do not contribute to the equations of motion as they are total derivatives, a non-minimal coupling of the form (5), (6) makes them no longer trivial.

IV 𝜸\gamma-Dual Effective Action

In this section we describe the construction of an effective action for inflation motivated by the γ\gamma-duality of the spinfoam dynamics. The γ\gamma-dual action SγS_{\gamma} is obtained via a duality rotation of a seed action SoddS_{\mathrm{odd}} that is parity-non-violating (odd under orientation-reversing diffeomorphisms \mathcal{R}). While, the action SoddS_{\mathrm{odd}} we start with here is chosen with the requirement that it provides a minimal model of inflation compatible with current observations [81], the construction of a γ\gamma-dual action via a duality rotation is more general and applies also to other possible seed actions.

We consider an action for gravity and a single scalar field with a potential that drives inflation, written in first order formalism:

Sodd(1)=12κΩEC+LmΩv+ΩS.S_{\mathrm{odd}}^{(1)}=\int\tfrac{1}{2\kappa}\Omega_{EC}\,+\,L_{m}\,\Omega_{\mathrm{v}}\,+\,\Omega_{S}\,. (61)

The action depends on eIe^{I} and ωIJ\omega^{IJ}, the Lorentz coframe and the Lorentz connection; and on ϕ\phi, and πI\pi^{I}, the scalar field and its covariant momentum. It is given by the integral of the Einstein-Cartan 44-form ΩEC\Omega_{EC} (36), the volume 44-form Ωv\Omega_{\mathrm{v}} (25), and the 44-form

ΩS=13!ϵIJKLπIdϕeJeKeL,\Omega_{S}=\frac{1}{3!}\epsilon_{IJKL}\,\pi^{I}\,d\phi\wedge e^{J}\wedge e^{K}\wedge e^{L}\,, (62)

that depends both on the scalar field ϕ\phi and its covariant momentum, the vector valued scalar πI\pi^{I}. The dynamics of the scalar field is encoded in the Lagrangian LmL_{m},

Lm=+12πIπIV(ϕ),L_{m}=+\tfrac{1}{2}\pi^{I}\pi_{I}-V(\phi)\,, (63)

with a potential V(ϕ)V(\phi) with a plateau that can drive slow-roll inflation [81], such as the Starobinsky potential [82]. This first-order action is equivalent to the one generally considered in single-field slow-roll inflation in the metric formalism: the stationarity of the action under variations of the Lorentz connection ωIJ\omega^{IJ} and of the scalar momentum πI\pi^{I} imposes that the torsion vanishes τI=eI=0\tau^{I}=\nabla e^{I}=0 and the momentum is given by πI=eIμμϕ\pi_{I}=e_{I}^{\mu}\partial_{\mu}\phi. Working in the Einstein-Cartan formalism, though, allows us to highlight three properties of this action that are relevant for our construction:

(i) The action Sodd(1)S_{\mathrm{odd}}^{(1)} is polynomial in the one-forms eIe^{I} and ωIJ\omega^{IJ}. In particular, it does not require that the coframe field eμI(x)dxμe^{I}_{\mu}(x)dx^{\mu} has an inverse eIμ=(eμI)1e^{\mu}_{I}=(e^{I}_{\mu})^{-1}. As such, it allows spinfoam-like 22d configurations of the geometry.

(ii) The action defines a theory that is invariant under spacetime diffeomorphisms Diff()\mathrm{Diff}(\mathcal{M}), including orientation-reversing diffeomorphisms \mathcal{R} as the action is parity-odd, Sodd(1)Sodd(1)S_{\mathrm{odd}}^{(1)}\overset{\mathcal{R}}{\longrightarrow}-S_{\mathrm{odd}}^{(1)}.

(iii) In an expansion in the number of exterior derivatives dd of fields, the term LmΩvL_{m}\,\Omega_{\mathrm{v}} is of zero-th order, while the two terms ΩEC\Omega_{EC} and ΩS\Omega_{S} are of first order as they depend on dωIJd\omega^{IJ} and dϕd\phi.

Following Weinberg’s construction of an effective field theory of inflation [20], we consider now an extension Sodd(2)S_{\mathrm{odd}}^{(2)} of the action Sodd(1)S_{\mathrm{odd}}^{(1)} that includes terms of second order in exterior derivatives, with the additional conditions that Sodd(2)S_{\mathrm{odd}}^{(2)} is (i) polynomial in the one-forms eIe^{I} and ωIJ\omega^{IJ}, and (ii) it is parity odd under orientation-reversing diffeomorphisms \mathcal{R}. These conditions exclude ΩTT\Omega_{TT} and ΩCS\Omega_{CS}, the torsion-squared and the Chern-Simons 44-forms which are parity-even (44), (46). We are left with the Gauss-Bonnet 44-form and Sodd(2)=f(ϕ)ΩGBS_{\mathrm{odd}}^{(2)}=\int f(\phi)\,\Omega_{GB}, where f(ϕ)f(\phi) is a function of the scalar field ϕ\phi. Note that we do not consider possible couplings to the covariant momentum πI\pi^{I} which effectively correspond to a modified kinetic term for the scalar field as in models of KK-inflation [83].

The parity non-violating action Sodd=Sodd(1)+Sodd(2)S_{\mathrm{odd}}=S_{\mathrm{odd}}^{(1)}+S_{\mathrm{odd}}^{(2)},

Sodd=12κΩEC+LmΩv+ΩS+f(ϕ)ΩGB,S_{\mathrm{odd}}=\int\tfrac{1}{2\kappa}\Omega_{EC}\,+\,L_{m}\,\Omega_{\mathrm{v}}\,+\,\Omega_{S}\,+\,f(\phi)\,\Omega_{GB}\,, (64)

is taken here as the starting point, or seed, for the construction of a γ\gamma-dual action for inflation. Under a duality rotation (4) of the Lorentz curvature FIJF^{IJ}, the 44-forms appearing in SoddS_{\mathrm{odd}} transform as444These transformation properties can equivalently be derived using the self-dual and antiself-dual variables FIJ(±)=12(FIJiFIJ)F^{(\pm)}_{IJ}=\frac{1}{2}(F_{IJ}\mp\mathrm{i}\,{}^{*}\!F_{IJ}) which transform as FIJ(±)𝜃e±iθFIJ(±)F^{(\pm)}_{IJ}\overset{\theta}{\longrightarrow}\mathrm{e}^{\pm\mathrm{i}\theta}F^{(\pm)}_{IJ}.

Ωv\displaystyle\Omega_{\mathrm{v}}\;\;\,\quad 𝜃Ωv,\displaystyle\overset{\theta}{\longrightarrow}\quad\Omega_{\mathrm{v}}\,, (65)
ΩS\displaystyle\Omega_{S}\;\;\quad 𝜃ΩS,\displaystyle\overset{\theta}{\longrightarrow}\quad\Omega_{S}\,, (66)
ΩEC\displaystyle\Omega_{EC}\quad 𝜃cosθΩEC+sinθΩH,\displaystyle\overset{\theta}{\longrightarrow}\quad\cos\theta\;\;\,\Omega_{EC}\;+\;\sin\theta\;\;\,\Omega_{H}\,, (67)
ΩGB\displaystyle\Omega_{GB}\quad 𝜃cos2θΩGB+sin2θΩCS,\displaystyle\overset{\theta}{\longrightarrow}\quad\cos 2\theta\;\Omega_{GB}\;+\;\sin 2\theta\;\Omega_{CS}\,, (68)

and define a one-parameter family of duality-rotated actions Sodd𝜃SθS_{\mathrm{odd}}\overset{\theta}{\longrightarrow}S_{\theta}.

We note that duality rotations of the curvature tensor have been considered multiple times in the literature with the different goal of relating solutions of Einstein equations [84, 85].555Note that here we consider only duality rotations in the Lorentz internal indices of FIJF^{IJ}. Alternatively, in the metric formalism, one can also consider a duality rotation of the Riemann tensor on spacetime manifold indices (via a tangent-space Hodge dual operator). These two definitions are not trivially equivalent; for a discussion on this comparison, see [86]. Recently, the duality rotation (67) of the Einstein-Cartan action has been considered in [17, 18] as an exact symmetry of vacuum Einstein gravity, but shown in [19] to be only a symmetry of a duality-preserving sector of vacuum solutions. The transformation considered here is not a symmetry of the action SoddS_{\mathrm{odd}}, but the construction of a one-parameter family of actions SθS_{\theta} that reproduces the relation between the EPRL spinfoam model and the parity-non-violating limit θ0\theta\to 0 given by the BC model discussed in Sec. II. By setting the parameter θ\theta to the value tanθ=1/γ\tan\theta=1/\gamma that relates it to the Barbero-Immirzi parameter γ\gamma (3) and normalizing the constant κ/cosθ=8πG\kappa/\cos\theta=8\pi G to the observed value of Newton’s constant GG as discussed in Sec. II, we obtain the spinfoam-motivated γ\gamma-dual action,

Sγ=Sγ(1)+Sγ(2)S_{\gamma}=S_{\gamma}^{(1)}+S_{\gamma}^{(2)} (69)

with

Sγ(1)[eI,ωIJ,ϕ]\displaystyle S_{\gamma}^{(1)}[e^{I},\omega^{IJ},\phi\,] =116πG(12ϵIJKLeIeJFKL1γeIeJFIJ)\displaystyle=\frac{1}{16\pi G}\!\int\left(\frac{1}{2}\epsilon_{IJKL}\,e^{I}\wedge e^{J}\wedge F^{KL}-\frac{1}{\gamma}\,e_{I}\wedge e_{J}\wedge F^{IJ}\right)
+(12ηIJeIμμϕeJννϕV(ϕ))14!ϵIJKLeIeJeKeL\displaystyle\quad+\int\left(-\frac{1}{2}\,\eta^{IJ}e^{\mu}_{I}\partial_{\mu}\phi\;e^{\nu}_{J}\partial_{\nu}\phi-V(\phi)\right)\frac{1}{4!}\,\epsilon_{IJKL}\,e^{I}\wedge e^{J}\wedge e^{K}\wedge e^{L}\, (70)
Sγ(2)[eI,ωIJ,ϕ]\displaystyle S_{\gamma}^{(2)}[e^{I},\omega^{IJ},\phi\,] =f(ϕ)(γ21γ2+112ϵIJKLFIJFKL+2γγ2+1FIJFIJ),\displaystyle=\int f(\phi)\left(\frac{\gamma^{2}-1}{\gamma^{2}+1}\;\frac{1}{2}\epsilon_{IJKL}\,F^{IJ}\wedge F^{KL}\,+\frac{2\gamma}{\gamma^{2}+1}F_{IJ}\wedge F^{IJ}\right)\,, (71)

where we have already solved for the covariant momentum πI\pi^{I}. Furthermore, imposing the zero-torsion condition τI=0\tau^{I}=0 as a constraint on the Lorentz connection,666Alternatively, one can implement a reduction of order procedure where one substitutes the solutions of the equations of motion of the first-order action Sγ(1)S_{\gamma}^{(1)} into the second order action Sγ(2)S_{\gamma}^{(2)} as discussed in [20]. we can write the γ\gamma-dual action in the metric formalism

Sγ[gμν,ϕ]=\displaystyle S_{\gamma}[g_{\mu\nu},\phi]= 116πGRgd4x+(12gμνμϕνϕV(ϕ))gd4x\displaystyle\;\;\frac{1}{16\pi G}\!\int R\,\sqrt{-g}\,\differential^{4}x\;\;+\int\Big{(}-\frac{1}{2}\,g^{\mu\nu}\partial_{\mu}\phi\,\partial_{\nu}\phi-V(\phi)\Big{)}\sqrt{-g}\,\differential^{4}x (72)
+f(ϕ)(γ212(γ2+1)(RμνρσRμνρσ4RμνRμν+R2)+γ2(γ2+1)ε(g)αβρσRμναβRμν)ρσgd4x.\displaystyle+\int f(\phi)\left(\frac{\gamma^{2}-1}{2(\gamma^{2}+1)}\left(R_{\mu\nu\rho\sigma}R^{\mu\nu\rho\sigma}-4R_{\mu\nu}R^{\mu\nu}+R^{2}\right)\,+\frac{\gamma}{2(\gamma^{2}+1)}\varepsilon_{(g)}^{\alpha\beta\rho\sigma}R_{\mu\nu\alpha\beta}R^{\mu\nu}{}_{\rho\sigma}\right)\sqrt{-g}\,\differential^{4}x\,.

In the remaining part of the paper we study primordial cosmological perturbations described by this parity-violating γ\gamma-dual action and the possibility of determining the Barbero-Immirzi parameter from observations.

Parity violating actions of the form (72) have been extensively studied [21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38]. Specifically, in [27, 28] an action for inflation that includes both the Gauss-Bonnet and the Chern-Simons terms was considered, but with independent couplings fGB(ϕ)f_{GB}(\phi) and fCS(ϕ)f_{CS}(\phi) to the scalar field. The requirement of γ\gamma-duality motivated by the spinfoam dynamics fixes the ratio of these two functions to a constant related to the Barbero-Immirzi parameter, fGB(ϕ)/fCS(ϕ)=γ1γf_{GB}(\phi)/f_{CS}(\phi)=\gamma-\frac{1}{\gamma}.

We note that, up to two conditions, the effective action (72) coincides with the action describing an effective field theory of inflation discussed in [20]. The logic there is to consider the most general action for a metric gμνg_{\mu\nu} and a scalar field ϕ\phi that is invariant under diffeomorphisms, it is organized in the number of derivatives (truncated at the fourth derivative), up to field redefinitions and with reduction of the order with respect to the truncation to second derivatives. Using reduction of order with respect to the equations of motion of the action Sγ(1)S_{\gamma}^{(1)}, allows one to write the Weyl-squared term CμνρσCμνρσC_{\mu\nu\rho\sigma}C^{\mu\nu\rho\sigma} as a Gauss-Bonnet term, up to a redefinition of coupling functions f(ϕ)f(\phi). The first difference with respect to [20] is that (72) does not include a KK-inflation type correction to the matter sector [83],

SK=fK(ϕ)(gμνμϕνϕ)2gd4x.S_{K}=\int f_{K}(\phi)\left(g^{\mu\nu}\partial_{\mu}\phi\,\partial_{\nu}\phi\right)^{2}\,\sqrt{-g}\,\differential^{4}x\,. (73)

In principle, this term can be included in our action but we do not explore it here. The second difference is that in [20] there are two independent coupling functions fGB(ϕ)f_{GB}(\phi) and fCS(ϕ)f_{CS}(\phi). Here, the new condition of γ\gamma-duality relates the ratio of the two to the Barbero-Immirzi parameter.

V Inflation and Primordial Gravitational Waves

In this section we study the power spectrum of primordial gravitational waves in a model of inflation given by the γ\gamma-dual action (72). We refer to [87, 88] for an introduction to cosmological perturbations and to [21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38] for the study of primordial gravitational waves from parity-violating actions including a Chern-Simons term.

We consider a homogeneous and isotropic, spatially flat (k=0)(k=0) Friedman-Lemaître-Robertson-Walker (FLRW) background. Tensorial perturbations are described by the perturbed metric

gμνdxμdxν=dt2+a(t)2(δij+2Eij(t,x))dxidxj,g_{\mu\nu}dx^{\mu}dx^{\nu}=-dt^{2}+a(t)^{2}\big{(}\delta_{ij}+2E_{ij}(t,\vec{x})\big{)}dx^{i}dx^{j}\,, (74)

where tt is cosmic time, a(t)a(t) the scale factor, the spatial indices i,j,i,j,\ldots are raised and lowered with the 3d3d Euclidean metric δij\delta_{ij}, and EijE_{ij} the transverse-traceless component in a scalar-vector-tensor decomposition of perturbations. In this section we do not consider scalar perturbations, which decouple from tensor perturbations in the quadratic expansion of the action we are working in.

A 3-dimensional Fourier transform allows us to work directly with the two circular polarizations (±)(\pm) per mode,

Eij(t,x)=d3k(2π)3eikxλ=±Jijλ(k)hλ(t,k),E_{ij}(t,\vec{x})=\int\frac{\differential^{3}\vec{k}}{(2\pi)^{3}}\,\mathrm{e}^{\mathrm{i}\vec{k}\cdot\vec{x}}\sum_{\lambda=\pm}J_{ij}^{\lambda}(\vec{k})\;h_{\lambda}\!\left(\text{\small{\hbox{t,\vec{k}}}}\right), (75)

where Jij±(k)J_{ij}^{\pm}(\vec{k}) is the circular polarization basis with kiJij±(k)=0k^{i}J_{ij}^{\pm}(\vec{k})=0, Jij+(k)=Jij(+k)J_{ij}^{+}(-\vec{k})=J_{ij}^{-}(+\vec{k}), and normalization Tr(J+(k)J(k))=4\mathrm{Tr}(J^{+}(\vec{k})J^{-}(\vec{k}))=4.

The equations of motion satisfied by the homogeneous and isotropic background solution a(t)a(t), ϕ(t)\phi(t) are

H2=κ3(ϕ˙22+V(ϕ)+ 12γ21γ2+1H3f˙(ϕ))\displaystyle H^{2}=\frac{\kappa}{3}\left(\frac{\dot{\phi}^{2}}{2}\,+\,V(\phi)\,+\,12\frac{\gamma^{2}-1}{\gamma^{2}+1}\,H^{3}\,\dot{f}(\phi)\right) (76)
ϕ¨+3Hϕ˙+V(ϕ)=12γ21γ2+1(H4+H2H˙)f(ϕ),\displaystyle\ddot{\phi}+3H\dot{\phi}+V^{\prime}(\phi)=12\frac{\gamma^{2}-1}{\gamma^{2}+1}\left(H^{4}+H^{2}\dot{H}\right)f^{\prime}(\phi), (77)

where H(t)=a˙(t)a(t)H(t)=\frac{\dot{a}(t)}{a(t)} is the Hubble rate, a dot (˙)(\,\dot{}\,) denotes cosmic time derivatives, and a prime ()(\,^{\prime}\,) denotes derivative with respect to the scalar field ϕ\phi.

To describe a slow-roll phase, one introduces the parameters

ϵV(t)12κ(V(ϕ)V(ϕ))2,ηV(t)1κV′′(ϕ)V(ϕ),\displaystyle\epsilon_{V}(t)\equiv\frac{1}{2\kappa}\left(\frac{V^{\prime}(\phi)}{V(\phi)}\right)^{2}\;,\quad\eta_{V}(t)\equiv\frac{1}{\kappa}\frac{V^{\prime\prime}(\phi)}{V(\phi)}\,, (78)
ϵH(t)H˙H2,δ(t)ϕ¨ϕ˙H,ρ2(t)κϕ˙2H2,\displaystyle\epsilon_{H}(t)\equiv-\frac{\dot{H}}{H^{2}}\,,\quad\delta(t)\equiv\frac{\ddot{\phi}}{\dot{\phi}H}\,,\quad\rho^{2}(t)\equiv\frac{\kappa\dot{\phi}^{2}}{H^{2}}\,, (79)

and demands that they are small. The introduction of a coupling f(ϕ)f(\phi) and higher-order corrections to the action requires us to define additional slow-roll parameters. Following [27, 28], we define

σ(t)\displaystyle\sigma(t) κf˙(ϕ)H(t)\displaystyle\equiv\kappa\dot{f}(\phi)H(t) (80)
α(t)\displaystyle\alpha(t) κ3V(ϕ)f(ϕ)\displaystyle\equiv\frac{\kappa}{3}V^{\prime}(\phi)f^{\prime}(\phi) (81)
β(t)\displaystyle\beta(t) κ3V(ϕ)f′′(ϕ)\displaystyle\equiv\frac{\kappa}{3}V(\phi)f^{\prime\prime}(\phi) (82)

and assume that they are small to ensure that f(ϕ)f(\phi) varies slowly. Notice that some parameters, e.g., σ,α,ηV\sigma,\,\alpha,\,\eta_{V}, can be negative.

All parameters are not just required to be small, but also to vary slowly in time. This implies a tower of slow-roll parameters, for each one of the quantities described above, defined as

σ2(t)σ˙(t)H(t)σ(t),σ3(t)σ˙2(t)H(t)σ2(t),\sigma_{2}(t)\equiv-\frac{\dot{\sigma}(t)}{H(t)\sigma(t)}\;,\;\;\sigma_{3}(t)\equiv-\frac{\dot{\sigma}_{2}(t)}{H(t)\sigma_{2}(t)}\;,\,\ldots (83)

and similarly for {ϵH,ϵH2,ϵH3,}\{\epsilon_{H},\epsilon_{H2},\epsilon_{H3},\ldots\}, etc., such that, all of them are small. We will collectively track the order of the slow-roll expansion with their maximum, ϵmax(ϵH(t),,|ηV(t)|,,|σ(t)|,|σ2(t)|,)1\epsilon\equiv\mathrm{max}(\epsilon_{H}(t),...,|\eta_{V}(t)|,...,|\sigma(t)|,|\sigma_{2}(t)|,...)\ll 1.

Note that the slow-roll parameters are not independent from each other, for example,

ϵH(t)\displaystyle\epsilon_{H}(t) =ϵV(t)2γ21γ2+1α(t)\displaystyle=\epsilon_{V}(t)-2\frac{\gamma^{2}-1}{\gamma^{2}+1}\alpha(t) (84)
δ(t)\displaystyle\delta(t) =ϵH(t)η(t)+4γ21γ2+1(β+2α)\displaystyle=\epsilon_{H}(t)-\eta(t)+4\frac{\gamma^{2}-1}{\gamma^{2}+1}(\beta+2\alpha) (85)
σ2(t)\displaystyle\sigma_{2}(t) =ϵH(t)(1+2β(t)α(t))+2β(t)γ21γ2+1δ(t)\displaystyle=\epsilon_{H}(t)\left(1+2\frac{\beta(t)}{\alpha(t)}\right)+2\beta(t)\frac{\gamma^{2}-1}{\gamma^{2}+1}-\delta(t) (86)

and of particular importance to us,

σ(t)=α(t)+2γ21γ2+1α2(t)ϵV(t).\sigma(t)=-\alpha(t)+2\frac{\gamma^{2}-1}{\gamma^{2}+1}\frac{\alpha^{2}(t)}{\epsilon_{V}(t)}. (87)

This relationship follows from the fact that the functional form of our coupling for the Gauss-Bonnet and the Chern-Simons terms has the same function f(ϕ)f(\phi). For other consistency conditions, see [27].

We can then write the background equations at the lowest order as

H2(t)\displaystyle H^{2}(t) =κ3V(ϕ)+𝒪(ϵ),\displaystyle=\frac{\kappa}{3}\,V(\phi)+\mathcal{O}(\epsilon)\,, (88)
ϕ˙(t)\displaystyle\dot{\phi}(t) =V(ϕ)3H(t)(14αϵVγ21γ2+1)+𝒪(ϵ).\displaystyle=-\frac{V^{\prime}(\phi)}{3H(t)}\left(1-\frac{4\alpha}{\epsilon_{V}}\frac{\gamma^{2}-1}{\gamma^{2}+1}\right)+\mathcal{O}(\epsilon). (89)

Here we see how there is an effective modification from f(ϕ)f(\phi) to the potential V(ϕ)V(\phi), with the appearance of the extra (constant to lowest order) α\alpha-term, coming from the Gauss-Bonnet modification to the action. In a specific inflationary model for V(ϕ)V(\phi), equation (89) can provide a constraint for σγ21γ2+1\sigma\frac{\gamma^{2}-1}{\gamma^{2}+1}: For example, a model with a coupling function such that f(ϕ)>0f^{\prime}(\phi)>0, and V(ϕ)>0V^{\prime}(\phi)>0, such as the Starobinsky potential [82],

VS(ϕ)=ΛS4(1e2κ3ϕ)2V_{S}(\phi)=\Lambda_{S}^{4}\left(1-\mathrm{e}^{-\sqrt{\frac{2\kappa}{3}}\phi}\right)^{2} (90)

would require

|σ|γ21γ2+1<ϵH4,\displaystyle|\sigma|\,\frac{\gamma^{2}-1}{\gamma^{2}+1}<\frac{\epsilon_{H}}{4}, (91)

in order for the field to consistently roll down the potential (ϕ˙<0\dot{\phi}<0).

The next step is to expand the action (72) around the slow-roll background, up to quadratic order in tensorial perturbations h±(t,k)h_{\pm}(t,\vec{k}). We find

Sγ±[h±]=12𝑑td3k(2π)3a3(Z±|h˙±|2k2a2B±|h±|2),S_{\gamma\pm}[h_{\pm}]=\frac{1}{2}\int\!dt\!\!\int\!\frac{\differential^{3}\vec{k}}{(2\pi)^{3}}\,a^{3}\Big{(}Z_{\pm}\,|\dot{h}_{\pm}|^{2}-\frac{k^{2}}{a^{2}}B_{\pm}|h_{\pm}|^{2}\Big{)}, (92)

with

Z±(t,k)=4κ(1+4κf˙(ϕ)(γ21γ2+1H(t)±2γγ2+1ka(t)))\displaystyle Z_{\pm}(t,k)=\textstyle\frac{4}{\kappa}\Big{(}1+4\kappa\dot{f}(\phi)\Big{(}\frac{\gamma^{2}-1}{\gamma^{2}+1}H(t)\pm\frac{2\gamma}{\gamma^{2}+1}\frac{k}{a(t)}\Big{)}\Big{)} (93)
B±(t,k)=4κ(1+4κ(γ21γ2+1f¨(ϕ)±2γγ2+1f˙(ϕ)ka(t)))\displaystyle B_{\pm}(t,k)=\textstyle\frac{4}{\kappa}\Big{(}1+4\kappa\Big{(}\frac{\gamma^{2}-1}{\gamma^{2}+1}\ddot{f}(\phi)\pm\frac{2\gamma}{\gamma^{2}+1}\dot{f}(\phi)\frac{k}{a(t)}\Big{)}\Big{)} (94)

and k=|k|k=|\vec{k}|. Note that the circular polarizations ±\pm have different quadratic actions and dynamics. A constant function f(ϕ)f(\phi) turns the correction null and recovers the action for tensorial perturbations in standard general relativity where Z±=B±=4/κZ_{\pm}=B_{\pm}=4/\kappa.

An expansion in the slow-roll parameters up to 𝒪(ϵ)\mathcal{O}(\epsilon) simplifies the functions (93) and (94) as second derivatives of the coupling function are of order 𝒪(ϵ2)\mathcal{O}(\epsilon^{2}),

Z±(t,k)=\displaystyle Z_{\pm}(t,k)= 4κ(1+4σ(t)(γ21γ2+1±2γγ2+1ka(t)H(t))),\displaystyle\textstyle\frac{4}{\kappa}\left(1+4\sigma(t)\!\left(\frac{\gamma^{2}-1}{\gamma^{2}+1}\pm\frac{2\gamma}{\gamma^{2}+1}\frac{k}{a(t)H(t)}\right)\right)\,, (95)
B±(t,k)=\displaystyle B_{\pm}(t,k)= 4κ(1±4σ(t)2γγ2+1ka(t)H(t)).\displaystyle\textstyle\frac{4}{\kappa}\left(1\pm 4\sigma(t)\,\frac{2\gamma}{\gamma^{2}+1}\frac{k}{a(t)H(t)}\right). (96)

We note that, as in Chern-Simons gravity [22, 23], here there is a perturbative instability because the functions Z±Z_{\pm} and B±B_{\pm} become negative at high momentum kk for one of the two circular polarizations [29]. However the kk-dependent terms in (95,96) appear together with the slow-roll parameter σ\sigma (80), making the band of stable modes sufficiently large. In contrast, in the general case where there are independent coupling functions fGB(ϕ)f_{GB}(\phi) and fCS(ϕ)f_{CS}(\phi) in (72), the background dynamics does not depend on the Chern-Simons coupling and the range of stable modes kk depends on a separate condition to be imposed on the function fCS(ϕ)f_{CS}(\phi) [27, 28].

We assume here that the value of γ\gamma is finite and not as small as the order of magnitude of the slow-roll parameters, or smaller, otherwise this expansion is not complete and we would have to consider corrections to the action of order ϵ2\epsilon^{2} to account for terms like σ2γγ2+1ka(t)H(t)\sigma\frac{2\gamma}{\gamma^{2}+1}\frac{k}{a(t)H(t)}. We will assume from now on that γ\gamma is not too small and the slow-roll approximation is valid at 𝒪(ϵ)\mathcal{O}(\epsilon).

The field operator h^±\hat{h}_{\pm} for the tensor perturbations of each circular polarization satisfies the equal-time commutation relations

[h^λ(t,k),ddth^λ(t,k)]=ia3Zλδλλ(2π)3δ3(kk)\Big{[}\hat{h}_{\lambda}(t,\vec{k}),\frac{d}{dt}\hat{h}_{\lambda^{\prime}}(t,-\vec{k}^{\prime})\Big{]}=\frac{\mathrm{i}\,\hbar}{a^{3}Z_{\lambda}}\delta_{\lambda\lambda^{\prime}}(2\pi)^{3}\delta^{3}(\vec{k}-\vec{k}^{\prime}) (97)

as follows from the canonical quantization of a theory with action (92) that includes a factor a3Zλa^{3}Z_{\lambda} in the kinetic term. A Fock representation of the field operators is given by the mode expansion

h^±(t,k)=u±(t,k)b^±(k)+u¯±(t,k)b^±(k),\hat{h}_{\pm}(t,\vec{k})=u_{\pm}(t,k)\,\hat{b}_{\pm}(\vec{k})+\bar{u}_{\pm}(t,k)\,\hat{b}^{\dagger}_{\pm}(-\vec{k})\,, (98)

where b^±\hat{b}_{\pm} are bosonic annihilation operators with [b^λ(k),b^λ(k)]=0[\hat{b}_{\lambda}(\vec{k}),\hat{b}_{\lambda^{\prime}}(\vec{k}^{\prime})]=0 and

[b^λ(k),b^λ(k)]=δλλ(2π)3δ3(kk),\displaystyle\Big{[}\hat{b}_{\lambda\vphantom{\lambda^{\prime}}}(\vec{k}),\hat{b}^{\dagger}_{\lambda^{\prime}}(\vec{k}^{\prime})\Big{]}=\delta_{\lambda\lambda^{\prime}}(2\pi)^{3}\delta^{3}(\vec{k}-\vec{k}^{\prime})\,, (99)

while u±(t,k)u_{\pm}(t,k) are mode functions (with u¯\bar{u} denoting the complex conjugate). We assume the mode functions to depend only on k=|k|k=|\vec{k}| to describe a Fock vacuum b^±(k)|0=0\hat{b}_{\pm}(\vec{k})|0\rangle=0 that has homogeneous and isotropic correlation functions. The mode functions u±(t,k)u_{\pm}(t,k) satisfy the canonical Wronskian conditions

u±(t,k)u¯˙±(t,k)u˙±(t,k)u¯±(t,k)=ia(t)3Z±(t,k),{u}_{\pm}(t,k)\dot{\bar{u}}_{\pm}(t,k)-\dot{u}_{\pm}(t,k)\bar{u}_{\pm}(t,k)=\frac{\mathrm{i}\,\hbar}{a(t)^{3}Z_{\pm}(t,k)}, (100)

that follows from the commutation relations (97), and the equations of motion

u¨±(t,k)+(3H(t)+Z˙±(t,k)Z±(t,k))u˙±(t,k)+k2a(t)2B±(t,k)Z±(t,k)u±(t,k)=0,\displaystyle\ddot{u}_{\pm}(t,k)+\left(3H(t)+\frac{\dot{Z}_{\pm}(t,k)}{Z_{\pm}(t,k)}\right)\dot{u}_{\pm}(t,k)+\frac{k^{2}}{a(t)^{2}}\frac{B_{\pm}(t,k)}{Z_{\pm}(t,k)}u_{\pm}(t,k)=0\,, (101)
u¨±(t,k)+3H(t)(143σ(t)2γγ2+1ka(t)H(t))u˙±(t,k)+k2a(t)2(14σ(t)γ21γ2+1)u±(t,k)=0,\displaystyle\ddot{u}_{\text{\tiny{$\pm$}}}(t,k)+3H(t)\left(1\mp\frac{4}{3}\sigma(t)\frac{2\gamma}{\gamma^{2}+1}\frac{k}{a(t)H(t)}\right)\dot{u}_{\text{\tiny{$\pm$}}}(t,k)+\frac{k^{2}}{a(t)^{2}}\left(1-4\sigma(t)\frac{\gamma^{2}-1}{\gamma^{2}+1}\right)u_{\text{\tiny{$\pm$}}}(t,k)=0\,, (102)

which we have written both in terms of the functions Z±Z_{\pm}, B±B_{\pm} and in a slow-roll expansion. A solution u±(t,k)u_{\pm}(t,k) of the equations (100, 101) that is ultraviolet adiabatic defines a Fock vacuum |0|0\rangle. Here we are interested in an adiabatic vacuum that generalizes the Bunch-Davies condition for modes in a band [kmin,kmax][k_{\min},k_{\max}] that includes the pivot mode kk_{\ast} at which to compute the amplitude and the tilt of the tensor power spectrum, with e.g. k=0.05Mpc1k_{\ast}=0.05\,\mathrm{Mpc}^{-1} for Planck satellite’s CMB measurements [81]. It is useful then to change the time parametrization to ts=ka(t)H(t)t\rightarrow s=\frac{k_{\ast}}{a(t)H(t)}. Denoting derivatives with respect to ss with a prime, the equation of motion (102) becomes,

uk,±′′(s)+(22ϵH(s)s±8γγ2+1σ(s)kk)uk,±(s)+k2k2(1+2ϵH(s)4γ21γ2+1σ(s))uk,±(s)=0.u_{k,\text{\tiny{{\hbox{\pm}}}}}^{\prime\prime}(s)+\left(\frac{-2-2\epsilon_{H}(s)}{s}\pm\frac{8\gamma}{\gamma^{2}+1}\sigma(s)\frac{k}{k_{\ast}}\right)u_{k,\text{\tiny{{\hbox{\pm}}}}}^{\prime}(s)+\frac{k^{2}}{k_{\ast}^{2}}\left(1+2\epsilon_{H}(s)-4\frac{\gamma^{2}-1}{\gamma^{2}+1}\sigma(s)\right)u_{k,\text{\tiny{{\hbox{\pm}}}}}(s)=0\,. (103)

We then change variables to

uk,±(s)yk,±(s)=Mk,±(s)uk,±(s),u_{k,\text{\tiny{{\hbox{\pm}}}}}(s)\rightarrow y_{k,\text{\tiny{{\hbox{\pm}}}}}(s)=\sqrt{M_{k,\text{\tiny{{\hbox{\pm}}}}}(s)}u_{k,\text{\tiny{{\hbox{\pm}}}}}(s), (104)

where

Mk,±(s)=k3(1ϵH(s))Zk,±(s)H(s)2s2,M_{k,\text{\tiny{{\hbox{\pm}}}}}(s)=\frac{k_{\ast}^{3}(1-\epsilon_{H}(s))Z_{k,\text{\tiny{{\hbox{\pm}}}}}(s)}{\hbar H(s)^{2}s^{2}}, (105)

after which, the equations of motion become

yk,±′′(s)+Qk,±(s)yk,±(s)=0.y_{k,\text{\tiny{{\hbox{\pm}}}}}^{\prime\prime}(s)+Q_{k,\text{\tiny{{\hbox{\pm}}}}}(s)\,y_{k,\text{\tiny{{\hbox{\pm}}}}}(s)=0. (106)

with

Qk,±(s)\displaystyle Q_{k,\text{\tiny{$\pm$}}}(s) =23ϵH(s)s2±8γγ2+1kkσ(s)s\displaystyle=\frac{-2-3\epsilon_{H}(s)}{s^{2}}\pm\frac{8\gamma}{\gamma^{2}+1}\frac{k}{k_{\ast}}\frac{\sigma(s)}{s}
+k2k2(1+2ϵH(s)4γ21γ2+1σ(s)),\displaystyle\qquad+\frac{k^{2}}{k_{\ast}^{2}}\left(1+2\epsilon_{H}(s)-4\frac{\gamma^{2}-1}{\gamma^{2}+1}\sigma(s)\right), (107)

and the canonical Wronskian condition becomes,

yk,±(s)y¯k,±(s)yk,±(s)y¯k,±(s)=i.y_{k,\text{\tiny{{\hbox{\pm}}}}}(s)\bar{y}^{\prime}_{k,\text{\tiny{{\hbox{\pm}}}}}(s)-y^{\prime}_{k,\text{\tiny{{\hbox{\pm}}}}}(s)\bar{y}_{k,\text{\tiny{{\hbox{\pm}}}}}(s)=\mathrm{i}. (108)

As a check, we note that the equation of motion (106) reduces to the standard equation for tensor modes in general relativity when σ0\sigma\rightarrow 0, i.e.,

yk′′(s)+(23ϵH(s)s2+(1+2ϵH(s))k2k2)yk(s)=0.y_{k}^{\prime\prime}(s)+\left(\frac{-2-3\epsilon_{H}(s)}{s^{2}}+(1+2\epsilon_{H}(s))\frac{k^{2}_{\ast}}{k^{2}}\right)y_{k}(s)=0. (109)

Since we are interested in a small band of modes around kk_{\ast}, and with modes freezing at s=1s=1, the slowly-varying Hubble rate and slow-roll parameters can themselves be expressed as a series on ss. Following [89], we introduce a log(s)\log(s) expansion for the Hubble rate and the slow-roll parameters,

H(s)=H(1+(ϵH+ϵH2+𝒪(ϵ3))log(s)+12(ϵH(ϵH+ϵH2)+ϵH2(2ϵH+3ϵH2)𝒪(ϵ4))log2(s)+),\displaystyle H(s)=H_{\ast}\,\big{(}1+(\epsilon_{H\ast}+\epsilon_{H\ast}^{2}+\mathcal{O}(\epsilon^{3}))\log(s)+\frac{1}{2}\left(\epsilon_{H\ast}(\epsilon_{H\ast}+\epsilon_{H2\ast})+\epsilon_{H\ast}^{2}(2\epsilon_{H\ast}+3\epsilon_{H2\ast})\mathcal{O}(\epsilon^{4})\right)\log^{2}(s)+...\big{)}\,, (110)
ϵH(s)=ϵH+(ϵHϵH2+𝒪(ϵ3))log(s)+(12ϵHϵH2(ϵH2+ϵH3)+𝒪(ϵ4))log2(s)+\displaystyle\epsilon_{H}(s)=\epsilon_{H\ast}+\left(\epsilon_{H\ast}\epsilon_{H2\ast}+\mathcal{O}(\epsilon^{3})\right)\log(s)+\Big{(}\frac{1}{2}\epsilon_{H\ast}\epsilon_{H2\ast}(\epsilon_{H2\ast}+\epsilon_{H3\ast})+\mathcal{O}(\epsilon^{4})\Big{)}\log^{2}(s)+... (111)

(where the asterisk denotes evaluation at s=1s=1 and the tracking parameter ϵ\epsilon is evaluated at freezing as well). For any slow-roll parameter, the coefficient for logn(s)\log^{n}(s) is at least of 𝒪(ϵn+1)\mathcal{O}(\epsilon^{n+1}); and for the Hubble rate, of 𝒪(ϵn)\mathcal{O}(\epsilon^{n}). We obtain then the equation of motion at 𝒪(ϵ)\mathcal{O}(\epsilon) around the freezing time of the mode kk_{\ast} to be of the form (106) and

Qk,±(s)=23ϵHs2±8γγ2+1kkσs+k2k2(1+2ϵH4γ21γ2+1σ)+𝒪(ϵ2).Q_{k,\text{\tiny{{\hbox{\pm}}}}}(s)=\frac{-2-3\epsilon_{H\ast}}{s^{2}}\pm\frac{8\gamma}{\gamma^{2}+1}\frac{k}{k_{\ast}}\frac{\sigma_{\ast}}{s}+\frac{k^{2}}{k_{\ast}^{2}}\left(1+2\epsilon_{H\ast}-4\frac{\gamma^{2}-1}{\gamma^{2}+1}\sigma_{\ast}\right)\;\;+\,\mathcal{O}(\epsilon^{2}). (112)

Similarly, the function M(s)M(s) takes the form

M(s)=4k3κH2s2(1ϵH(1+2log(s))+4σ(γ21γ2+1±2γγ2+1kks))+𝒪(ϵ2).M(s)=\frac{4k_{\ast}^{3}}{\hbar\kappa\,H_{\ast}^{2}s^{2}}\left(1-\epsilon_{H\ast}\left(1+2\log(s)\right)+4\sigma_{\ast}\left(\frac{\gamma^{2}-1}{\gamma^{2}+1}\pm\frac{2\gamma}{\gamma^{2}+1}\frac{k}{k_{\ast}}s\right)\right)\;\;+\,\mathcal{O}(\epsilon^{2}). (113)

This equation of motion has the Whittaker functions as a basis of solutions, {y1,±(s),y2,±(s)}\{y_{1,\text{\tiny{$\pm$}}}(s),y_{2,\text{\tiny{$\pm$}}}(s)\},

y1,±(s)\displaystyle y_{1,\text{\tiny{$\pm$}}}(s) =αiμ0,ν0(iκ0s)+β𝒲iμ0,ν0(iκ0s)\displaystyle=\alpha\,\mathcal{M}_{\mp\mathrm{i}\mu_{0},\nu_{0}}\left(\mathrm{i}\kappa_{0}s\right)\,+\,\beta\,\mathcal{W}_{\mp\mathrm{i}\mu_{0},\nu_{0}}\left(\mathrm{i}\kappa_{0}s\right) (114)
y2,±(s)\displaystyle y_{2,\text{\tiny{$\pm$}}}(s) =y¯1,±(s),\displaystyle=\bar{y}_{1,\text{\tiny{$\pm$}}}(s), (115)

with

μ0\displaystyle\mu_{0} =4γγ2+1σ+𝒪(ϵ2),\displaystyle=\frac{4\,\gamma}{\gamma^{2}+1}\sigma_{\ast}+\mathcal{O}(\epsilon^{2})\,, (116)
ν0\displaystyle\nu_{0} =32+ϵH+𝒪(ϵ2),\displaystyle=\frac{3}{2}+\epsilon_{H\ast}+\mathcal{O}(\epsilon^{2})\,, (117)
κ0\displaystyle\kappa_{0} =2kk(1+ϵH2γ21γ2+1σ)+𝒪(ϵ2).\displaystyle=2\,\frac{k}{k_{\ast}}\left(1+\epsilon_{H\ast}-2\frac{\gamma^{2}-1}{\gamma^{2}+1}\sigma_{\ast}\right)+\mathcal{O}(\epsilon^{2})\,. (118)

The coefficients α\alpha and β\beta are fully determined by the canonical Wronskian condition (108) and a choice of initial adiabatic Bunch-Davies-like vacuum state; i.e., such that its two-point vacuum correlation function matches the one of Minkowski on the infinite past,

limsy¯±(s)y±(s)=k2k,\displaystyle\lim_{s\rightarrow\infty}\bar{y}_{\pm}(s)y_{\pm}(s)=\frac{{k_{\ast}}}{2k}\,, (119)
limsy±(s)=k2keikks.\displaystyle\lim_{s\rightarrow\infty}y_{\pm}(s)=\sqrt{\frac{{k_{\ast}}}{2k}}\mathrm{e}^{\mathrm{i}\frac{k}{k_{\ast}}s}. (120)

These conditions determine the adiabatic solutions given by

yad,±(s)=eπ2μ0κ0𝒲iμ0,ν0(iκ0s).\displaystyle{y_{\mathrm{ad},\text{\tiny{$\pm$}}}}(s)=\frac{\mathrm{e}^{\mp\frac{\pi}{2}\mu_{0}}}{\sqrt{\kappa_{0}}}\,\mathcal{W}_{\mp\mathrm{i}\mu_{0},\nu_{0}}\left(\mathrm{i}\kappa_{0}s\right)\,. (121)

Using the asymptotic behavior of the Whittaker 𝒲\mathcal{W} function, and defining yasy,±(s)yad,±(s){y_{\mathrm{asy},\text{\tiny{$\pm$}}}}(s)\sim{y_{\mathrm{ad},\text{\tiny{$\pm$}}}}(s) in the s0s\to 0 limit, we have

yasy,±(s)=eπ4(i2μ0)sν0((iκ0)ν0Γ(2ν0)Γ(12+ν0±iμ0)s12+𝒪(s32)),{y_{\mathrm{asy},\text{\tiny{{\hbox{\pm}}}}}}(s)=\mathrm{e}^{\frac{\pi}{4}(\mathrm{i}\mp 2\mu_{0})}s^{-\nu_{0}}\left(\frac{(\mathrm{i}\kappa_{0})^{-\nu_{0}}\Gamma(2\nu_{0})}{\Gamma(\text{\small{\hbox{\frac{1}{2}}}}+\nu_{0}\pm\mathrm{i}\mu_{0})}s^{\frac{1}{2}}+\mathcal{O}(s^{\frac{3}{2}})\right), (122)

where Γ\Gamma is the Gamma function. With this expression and the function (113), we can compute the power spectrum Δ±2(k)\Delta^{2}_{\pm}(k) of circularly polarized tensor perturbations,

Δ±2(k)\displaystyle\Delta^{2}_{\pm}(k) limtk32π2|u±(t,k)|2=lims0k32π2|yasy,±(s)|2M(s)\displaystyle\equiv\lim_{t\to\infty}\frac{k^{3}}{2\pi^{2}}\left|u_{\pm}(t,k)\right|^{2}=\lim_{s\rightarrow 0}\frac{k^{3}}{2\pi^{2}}\frac{\left|{y_{\mathrm{asy},\text{\tiny{$\pm$}}}}(s)\right|^{2}}{M(s)}
=GH22π(1+2ϵH(1ΓElog(2kk))\displaystyle=\frac{G\hbar\,H_{\ast}^{2}}{2\pi}\left(1+2\epsilon_{H\ast}\left(1-\Gamma_{E}-\log\left(\frac{2k}{k_{\ast}}\right)\right)\right.
+2σ(12(1±πγ)γ2+1))+𝒪(ϵ2),\displaystyle\left.\qquad+2\sigma_{\ast}\left(1-\frac{2(1\pm\pi\gamma)}{\gamma^{2}+1}\right)\right)\,+\,\mathcal{O}(\epsilon^{2}), (123)

where ΓE0.577\Gamma_{E}\approx 0.577 is Euler’s constant. From this expression, we can read, at first order in the slow-roll parameters, the amplitude at the pivot mode kk_{\ast},

A±\displaystyle A_{\ast\pm} Δ±2(k)=GNH22π(1+2ϵH(1ΓElog2)\displaystyle\equiv\Delta^{2}_{\pm}(k_{\ast})=\frac{G_{N}\hbar H_{\ast}^{2}}{2\pi}\left(1+2\epsilon_{H\ast}\left(1-\Gamma_{E}-\log 2\right)\right.
+2σ(12(1±πγ)γ2+1)),\displaystyle\left.\qquad\qquad\quad\quad+2\sigma_{\ast}\left(1-\frac{2(1\pm\pi\gamma)}{\gamma^{2}+1}\right)\right)\,, (124)

and the tensor tilt at the pivot mode,

nTdlog(Δ±2(k))dlog(k)|k=2ϵH.n_{T\ast}\equiv\left.\frac{d\log(\Delta^{2}_{\pm}(k))}{d\log(k)}\right|_{k_{\ast}}=-2\epsilon_{H\ast}\,. (125)

While the amplitude depends on the circular polarization of the modes, at this order in the slow-roll expansion, the tilt does not. Finally, the polarization of the mode kk_{\ast} is

ΠA+AA++A=4πσγγ2+1.\Pi_{\ast}\equiv\frac{A_{\ast+}-A_{\ast-}}{A_{\ast+}+A_{\ast-}}=-4\pi\sigma_{\ast}\frac{\gamma}{\gamma^{2}+1}. (126)

The polarization is non-vanishing for σ0\sigma_{\ast}\neq 0 and depends on the value of the Barbero-Immirzi parameter γ\gamma, see Fig. 1.

The expressions for the tilt nTn_{T\ast} and the polarization Π\Pi_{\ast} (125,126) are in agreement with the ones found in a more general case in [27, 28] where a different approximation for the mode equation is used. In particular the log(s)\log(s) expansion (110) allows us to determine the late time limit (V) in which the power spectrum is explicitly time independent.

VI Discussion

Refer to caption
Refer to caption
Figure 1: (Left.) Dependence of the polarization of primordial gravitational waves at the pivot mode kk_{\ast} on the Barbero-Immirzi parameter for the γ\gamma-dual action (72), assuming different values of the slow-roll parameter σ\sigma_{\ast}, (80). (Right.) Value of γ\gamma determined by a possible measurement of the cosmological observable qq, defined in terms of the tensor polarization, tensor tilt and tensor-to-scalar ratio (131). For reference we report the values γ00.127\gamma_{0}\simeq 0.127 (dashed line) and γM0.238\gamma_{M}\simeq 0.238 (dotted line) considered in [51] and [52].

In this paper we have introduced a new condition on parity-violating effective actions, the notion of γ\gamma-duality, that is motivated by the specific dependence of the spinfoam dynamics WγW_{\gamma} on this parameter. While at present there is no top-down derivation of an effective field theory from the full non-perturbative spinfoam dynamics of loop quantum gravity, the γ\gamma-dual effective action (72) encodes in a crucial way an exact property of the non-perturbative theory: spacetime diffeomorphisms Diff0()\mathrm{Diff}_{0}(\mathcal{M}) that are connected to the identity are an exact gauge symmetry, while the invariance under large diffeomorphisms such as orientation reversals \mathcal{R} is broken in a way that is controlled by the single parameter γ\gamma.

The possibility of determining the Barbero-Immirzi parameter γ\gamma from observations was first considered in [26] as a measure of parity violation in primordial gravitational waves, and then in [47, 48] as a parity-violating coupling of fermions to gravity. This condition of γ\gamma-duality relates the coupling constants in the effective action and therefore allows us to probe the dependence of observables on the parameter γ\gamma at the level of higher-curvature terms.

As an effective theory for cosmological inflation, the γ\gamma-dual action (72) predicts a polarization for primordial gravitational waves given by the relation (126),

ΠA+AA++A=4πσγγ2+1.\Pi\equiv\frac{A_{+}-A_{-}}{A_{+}+A_{-}}=-4\pi\sigma_{\ast}\frac{\gamma}{\gamma^{2}+1}. (127)

The difference in amplitude between the two circular polarizations is a direct consequence of the parity-violating nature of the action [21, 22, 27, 28]. Besides the dependence on γ\gamma, we note that the polarization Π\Pi depends also on the value of the slow-roll parameter σ\sigma_{\ast} at the pivot mode kk_{\ast}, which measures the shape of the coupling function f(ϕ)f(\phi) as defined in (80), (see Fig. 1, Left). Therefore, determining γ\gamma requires an independent measurement of other observables besides the tensor polarization Π\Pi. A similar dependence on additional parameters besides γ\gamma appears in the analysis of parity-violating observables in the coupling of fermions to gravity [47, 48].

The scalar power spectrum in an effective theory of the class considered here was derived in [27, 28], assuming general coupling functions fGB(ϕ)f_{GB}(\phi) and fCS(ϕ)f_{CS}(\phi) to the Gauss-Bonnet and Chern-Simons terms. Expressing the result of [27, 28] in terms the coupling functions of the γ\gamma-dual action (72), we find that the tensor-to-scalar ratio at the pivot mode kk_{\ast} is

r=16ϵH+32γ21γ2+1σ,r=16\,\epsilon_{H\ast}+32\frac{\gamma^{2}-1}{\gamma^{2}+1}\sigma_{\ast}\,, (128)

which depends on the slow-roll parameters ϵH\epsilon_{H\ast} and σ\sigma_{\ast} at the pivot mode, besides γ\gamma. Moreover, the tensor tilt is given by the expression (125),

nT=2ϵH.n_{T}=-2\epsilon_{H\ast}\,. (129)

Combining the primordial cosmological observables Π\Pi, rr and nTn_{T}, we can extract a relation that depends only on the Barbero-Immirzi parameter γ\gamma,

1γγ=π88nT+rΠ.\frac{1}{\gamma}-\gamma=\frac{\pi}{8}\frac{8n_{T}+r}{\Pi}. (130)

The relation (130) can be solved for positive value of the Barbero-Immirzi parameter in terms of the ratio qq, (see Fig. 1, Right):

γ=1+(q2)2q2,withqπ88nT+rΠ.\gamma\,=\,\sqrt{1+\left(\frac{q}{2}\right)^{2}\,}-\frac{q}{2}\,,\quad\textrm{with}\;\;q\equiv\frac{\pi}{8}\frac{8n_{T}+r}{\Pi}. (131)

In this way, a measurement of the polarization and tilt of primordial gravitational waves, together with their relative amplitude to the scalar modes, can provide a measurement of the Barbero-Immirzi parameter and, therefore, of the scale of the discreteness (1) in loop quantum gravity.

The relation (131) arises from the analysis of slow-roll inflation with the γ\gamma-dual effective action (72). We expect a similar relation to arise when one starts from a more general parity-non-violating action that includes also KK-inflation terms and is then duality-rotated to a γ\gamma-dual action. It is then interesting to investigate the possibility of determining γ\gamma in such theories using future cosmological observations of the polarization Π\Pi that appears in the denominator in (131), and of the consistency relation r+8nTr+8n_{T} that appears in the numerator:

(a) Polarization. There are prospects of measurements of the polarization Π\Pi in the near future with Cosmic Microwave Background (CMB) polarization experiments, including BICEP2/Keck [90], SPT [91], SPIDER [92], LiteBIRD [93], and CMB-S4 [94] experiments. For a comprehensive review on the main challenges and strategies for extracting the contribution from primordial gravitational waves to the BB-modes in the CMB, see [95]. While scalar perturbations can only imprint (even) E-polarization modes on the CMB, tensor perturbations can induce both (even) EE and (odd) BB-polarization modes [96, 97]. Furthermore, a measurement of non-zero two-point correlations between the temperature and BB-modes of the CMB, ClTB0C_{l}^{TB}\neq 0, or between the EE and BB-modes, ClEB0C_{l}^{EB}\neq 0, would be a definite indication of a parity violating phenomenon [96]. As shown in [26], the contribution from tensor modes to ClEBC_{l}^{EB} and ClTEC_{l}^{TE} go respectively as Δ+2(k)Δ2(k)\Delta^{2}_{+}(k)-\Delta^{2}_{-}(k) (polarization) and Δ+2(k)+Δ2(k)\Delta^{2}_{+}(k)+\Delta^{2}_{-}(k) (average amplitude), convoluted with the corresponding radiation transfer functions; and therefore, their measurement could be used to estimate the tensor polarization. For instance, in [98] one can find an analysis of the constraints on tensor polarization from the EBEB and TBTB correlations and the prospects for satellite-like missions for models with parity violation and a polarization Π\Pi.

(b) Consistency relations. CMB-S4 [94] experiments are expected to measure, or put stronger bounds on, the tensor-to-scalar ratio rr in the near future. Together with a measurement of the tensor tilt nTn_{T}, these observations will provide a probe of the consistency relation r=8nTr=-8n_{T}, a property of single-field slow-roll inflation in pure General Relativity. The effective theory described by the action (72) violates the consistency relations and the quantity r+8nTr+8n_{T} appears in the numerator in (130). While, in principle, we can have |r+8nT|1|r+8n_{T}|\ll 1 and |Π|1|\Pi|\ll 1, which would make observations difficult, the value of γ\gamma depends only on their ratio. For |r+8nT||Π||r+8n_{T}|\ll|\Pi|, we find a Barbero-Immirzi parameter γ1\gamma\approx 1, while for |r+8nT||Π||r+8n_{T}|\gg|\Pi| we have γ1\gamma\ll 1.

Here we have focused exclusively on the primordial power spectrum. It would be interesting to extend this analysis of bounds on the Barbero-Immirzi parameter via γ\gamma-dual effective actions for inflation, to other cosmological observables such as parity-violating features in primordial non-Gaussianities [30, 31, 32, 33, 34, 35, 36, 37, 38], and their possible imprint in the large scale structure of the Universe [99, 100, 101] and galaxy shapes [102].


Acknowledgments. We thank Pietro Donà, Juan Margalef-Bentabol, Djordje Minic, Javier Olmedo, Marc Schneider for useful discussions. This work was made possible through the support of the ID# 62312 grant from the John Templeton Foundation, as part of the project “The Quantum Information Structure of Spacetime” (QISS). The opinions expressed in this work are those of the author(s) and do not necessarily reflect the views of the John Templeton Foundation. E.B. acknowledges support from the National Science Foundation, Grant No. PHY-2207851.

References