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

Circumventing the Ricci-inverse no-go theorem with
complexifiable singularities: a novel dark energy model
In memory of Prof. A.Vitturi, for his kind and always helpful soul.

Mattia Scomparin [email protected] Mogliano Veneto, 31021, Italy
Abstract

Ricci-inverse gravity is a new type of fourth-order gravity theory based on the anti-curvature tensor, that is, the inverse of the Ricci tensor. In this context, we introduce a novel method to circumvent the binding effects of a well-known no-go theorem for cosmic trajectories that cannot smoothly join a decelerated cosmic age with the current accelerated expansion of the universe. We therefore design a new class of Ricci-inverse theories whose cosmologies, without falling into no-go singularities, achieve the observed expansion as a stable attractor solution. This new perspective retrains Ricci-inverse cosmologies as viable dark energy models.

I Introduction

Many cosmological observations require a modification to Einstein’s theory of General Relativity (GR) to explain the late-time accelerated expansion Riess et al. (1998); Abbott et al. (2019); Scolnic et al. (2018a) and the early-time cosmological inflation of our universe Starobinsky (1980); Guth (1981a); Hinshaw et al. (2013); Malik et al. (2023a).

There are many ways to construct alternative theories to GR. A possible approach in this direction consists of constructing appropriate modified gravity (MG) theories, in which an alteration of the well-known Einstein Hilbert (EH) action is considered Nojiri et al. (2017). On the other hand, most of these reformulations require troublesome assumptions in their foundations, such as the definition of dimensional coupling constants to be finely-tuned or the introduction of extra dynamical fields that have no obvious relation with Einstein’s strictly geometric formulation of GR.

Motivated by the above considerations, Amendola et al. proposed in Ref. Amendola et al. (2020) a MG theory called Ricci-inverse gravity whose gravitational Lagrangian density g(R,A)\mathcal{L}_{g}(R,A) is constructed by combining the Ricci scalar RR with a new purely geometrical object called anti-curvature scalar A=gμνAμνA=g_{\mu\nu}A^{\mu\nu}. By construction, the anti-curvature scalar is the trace of anti-curvature tensor Aμν=Rμν1A^{\mu\nu}=R_{\mu\nu}^{-1}, which is the inverse of the Ricci tensor RμνR_{\mu\nu} defined by the relation

AμσRσν=δνμ.A^{\mu\sigma}R_{\sigma\nu}=\delta^{\mu}_{\nu}\,. (1)

Here, gμνg_{\mu\nu} is the metric tensor and δνμ\delta^{\mu}_{\nu} is the Kronecker delta function. For the avoidance of doubt, one must pay attention to the fact that AR1A\neq R^{-1}. However, focusing only on the physical dimension, [A]=[R]1[A]=[R]^{-1}.

As a consequence, one can use the anti-curvature scalar to introduce new Lagrangian terms into EH action which, resulting directly from RR, share its same physical dimension and do not invoke any new dynamical field to modify GR. Among all scale-free possibilities, we mention: (i) the pure anti-curvature term A1A^{-1}, (ii) the general power-law profile AR+1A^{\ell}R^{\ell+1}, and (iii) any general profile of the form g(RA)Rg(RA)R. Note that (i) belongs to (ii) when =1\ell=-1, while (iii) reduces to (ii) when g(RA)=(RA)g(RA)=(RA)^{\ell}.

In recent years, Ricci-inverse gravity has been extensively studied in the currently available literature. Examples of applications include the investigation of: (i) anisotropic Shamir et al. (2023), compact Shamir and Meer (2023) and charged Malik et al. (2023b, 2024a) star structures Malik et al. (2024b, 2023a), (ii) matter-antimatter asymmetry phenomena Jawad and Sultan (2022), (iii) novel aspects related to wormhole Mustafa et al. (2024) and blackholes Ahmed and Guvendi (2024) solutions, (iv) axially symmetric spacetimes with causality violation de Souza and Santos (2023); de Souza et al. (2024), and (v) cosmic structures within a Sub-Horizon non-relativistic Weak-Field limit Scomparin (2021). Although most of these works emphasize the increasing relevance and applicability of Ricci-inverse gravity in diverse scenarios, two binding no-go theorems have been shown to rule out many cosmological realizations of this theory.

The first no-go theorem concerns the late-time accelerated expansion phase of universe. Its original formulation states that any Lagrangian density g(R,A)\mathcal{L}_{g}(R,A) containing terms proportional to AA^{\ell}, with any positive or negative \ell, cannot smoothly join a cosmic decelerated era with the current accelerated expansion of universe Aghanim et al. (2020) (see Refs.Amendola et al. (2020); Das et al. (2022) for a detailed discussion). In fact, assuming a standard Friedmann-Lemaître-Robertson-Walker (FLRW) metric, between these two epochs AA must pass through both 0 and ±\pm\infty values, leading a generic power of AA to blow up due to singularities. Up to now, attempts to find loopholes to escape this no-go theorem have not proved satisfactory. Indeed, it has been shown that the shiftings introduced by spatially curved metrics or anisotropic backgrounds are too small to move the singularities sourced by AA outside the observational range. In the end, only well-though non-polynomial Lagrangians (i.e. respecting the property [𝒫1][\mathcal{P}1] of being regular both for A0A\rightarrow 0 and A±A\rightarrow\pm\infty) seem promising enough to circumvent the problem, but they are very complicated and consequently have not been investigated so far Amendola et al. (2020); Das et al. (2022).

Another cosmic scenario now forbidden to Ricci-inverse gravity is the so-called inflationary era, which occurred very rapidly after the Big Bang Starobinsky (1980); Guth (1981b); Linde (1982). Although this primordial phase is not affected by singularities from anti-curvature terms, a stability analysis based on a dynamical system method showed that, even when g(R,A)\mathcal{L}_{g}(R,A) depends on simple powers of AA (e.g AA or A2A^{2}), all isotropic inflationary solutions turn out to be unstable with respect to field perturbations Do (2021, 2022). This behavior is due to the fact that the existence domain and the stability region of the solutions do not overlap. Therefore, the impossibility to have a stable (isotropic) inflationary solutions as well as a smooth exit from inflation era (see, e.g. Ref. Das et al. (2022)) in elementary Ricci-inverse cosmologies is essentially the content of the second no-go theorem.

In the absence of effective loopholes, all these results have raised great doubts about the cosmological viability of Ricci-inverse gravity to be a suitable inflationary model or DE candidate. To the author’s knowledge, no valid proposal has been found at present to circumvent the action of the two no-go theorems. Furthermore, neither non-trivial extensions involving higher-order scalar combinations in the anti-curvature tensor111Here, AναργA_{\nu}^{\alpha\rho\gamma} is the anti-Riemann tensor, defined by the identity AναργRαρβμ=δνμδβγA^{\alpha\rho\gamma}_{\nu}R^{\mu}_{\alpha\rho\beta}=\delta^{\mu}_{\nu}\delta^{\gamma}_{\beta}. such as g(R,AμνAμν)\mathcal{L}_{g}(R,A^{\mu\nu}A_{\mu\nu}) or g(R,AμαβγAαβγμ)\mathcal{L}_{g}(R,A^{\alpha\beta\gamma}_{\mu}A_{\alpha\beta\gamma}^{\mu}) seem a promising prospect for treating in a simple way the pathologies that afflict the Ricci-inverse gravity, which, in the end, is still ruled out Das et al. (2022).

At this point, one might come to the conclusion that there is no way around no-go theorems for Ricci-inverse cosmologies. Of course, every no-go theorem is based on arguments that one can try to break or reconsider. For example, in the case of the first no-go theorem, the most suitable point to be reconsidered is that 𝒫1\mathcal{P}1 is really the only property that makes non-polynomial Lagrangians capable of realizing regular equations of motion. In fact, looking for a possible workaround of the theorem in this way, the author realized that there is indeed a new property that can regularize the no-go singularities.

Generally speaking, our reasoning is as follows. Focusing on Ricci-inverse cosmologies studied so far, we noticed that their Lagrangians exhibit no-go singularities belonging only to the real axis \mathbb{R}. However, if such a Lagrangians could be redrawn to push the no-go singularities into the complex plane \mathbb{C}, we might of course expect the effects of the first no-go theorem to become inoffensive, since the cosmic trajectories are and remain purely real. We call this property [𝒫2][\mathcal{P}2], i.e. regularizability with complexifiable singularities. It refers to the characteristic of some non-polynomial Lagrangians, under certain conditions on their free parameters, to shift their no-go singularities to the complex plane with a non-zero imaginary part.

Driven by this idea, the aim of this paper is twofold: (i) to review the known aspects about the first no-go theorem in Ricci-inverse gravity in the form of g(R,A)\mathcal{L}_{g}(R,A), and (ii) to present new original results on this topic. In particular, using [𝒫2][\mathcal{P}2], we present a novel singularity-free DE model whose cosmic trajectories smoothly connect a cosmic decelerated era with the current accelerated expansion of the universe, which in turn emerges as a stable attractor solution alongside a de Sitter phase.

The outline of our work is as follows. In Section II we begin by briefly introducing the full Ricci-inverse theory g(R,A)\mathcal{L}_{g}(R,A) and discussing the related covariant field equations. In Section III we introduce the Ricci and anti-curvature scalars in a flat-space FLRW universe and, contextually, in Subsection III.1 we review the state of the art on the first no-go theorem. Section IV is devoted to the introduction of a novel class of Ricci-inverse models that might be expected to violate the first no-go theorem. Such models depend on two free-parameters and in Subsection IV.1 we explain how to implement our property [𝒫2][\mathcal{P}2] to tune them and obtain healthy Ricci-inverse cosmological solutions. After finding the related modified Friedmann equations in Section V, in Subsection V.1 and Subsection V.2 we provide proofs of stable de Sitter and power-law solutions (compatible with the current accelerated phase). Finally, in Section VI we summarise our results and identify future avenues of investigation..

In this work, we use the metric signature (,+,+,+)(-,+,+,+) and Einstein’s convention on repeated dummy indices is assumed.

II Ricci-inverse gravitational theory

Let us consider the full Ricci-inverse theory described by the basic action

S=(12κg+m)g𝑑td3x,S=\int\!\left(\frac{1}{2\kappa}\mathcal{L}_{g}+\mathcal{L}_{m}\right)\sqrt{-g}\,dtd^{3}\!x\,, (2)

where gg is the determinant of the metric gμνg_{\mu\nu} and κ8πG/c4\kappa\equiv 8\pi G/c^{4} is a constant that depends on the gravitational parameter GG and the speed of light cc. The first term g=g(R,A)\mathcal{L}_{g}=\mathcal{L}_{g}(R,A) corresponds to the MG Lagrangian density, which is an a priori arbitrary function of the Ricci and anti-curvature scalars. The second term m\mathcal{L}_{m} is the matter Lagrangian density, that we assume minimally coupled with the metric tensor only.

It is immediate to check that if g=R\mathcal{L}_{g}=R, then Eq. (2) reduces to the usual EH action for GR.

Varying the above action with respect to the metric, the field equation is δS=0\delta S=0 Amendola et al. (2020); Lee (2021). By differentiating definition (1) and using integration by parts we obtain

0=δSδgμνδgμν=12g[𝒢μνκTμν]δgμν𝑑td3x,0=\frac{\delta S}{\delta g^{\mu\nu}}\delta g_{\mu\nu}=\frac{1}{2}\int\!\sqrt{-g}\,\big{[}\mathcal{G}^{\mu\nu}-\kappa T^{\mu\nu}\big{]}\delta g_{\mu\nu}\,dtd^{3}\!x\,, (3)

where we introduced the modified Einstein tensor

𝒢μν\displaystyle\mathcal{G}^{\mu\nu} RgRμν12ggμν\displaystyle\equiv\partial_{R}\mathcal{L}_{g}R^{\mu\nu}-\tfrac{1}{2}\mathcal{L}_{g}g^{\mu\nu}
AgAμν\displaystyle-\partial_{A}\mathcal{L}_{g}A^{\mu\nu}
12αα(AgAσμAνσ)\displaystyle-\tfrac{1}{2}\nabla^{\alpha}{\nabla_{\alpha}{(\partial_{A}\mathcal{L}_{g}A_{\sigma}^{\mu}A^{\nu\sigma}})} (4)
+gρμ(αρAg)AσαAνσ\displaystyle+g^{\rho\mu}(\nabla_{\alpha}{\nabla_{\rho}{\,\partial_{A}\mathcal{L}_{g}}})A^{\alpha}_{\sigma}A^{\nu\sigma}
12gμναβ(AgAσαAβσ)\displaystyle-\tfrac{1}{2}g^{\mu\nu}\nabla_{\alpha}{\nabla_{\beta}{(\partial_{A}\mathcal{L}_{g}A_{\sigma}^{\alpha}A^{\beta\sigma})}}
μνRg+gμνααRg,\displaystyle-\nabla^{\mu}{\nabla^{\nu}{\partial_{R}\mathcal{L}_{g}}}+g^{\mu\nu}\nabla^{\alpha}{\nabla_{\alpha}{\,\partial_{R}\mathcal{L}_{g}}}\,,

and the matter energy-momentum tensor

Tμν2gδ(gm)δgμν.T^{\mu\nu}\equiv\frac{2}{\sqrt{-g}}\frac{\delta(\sqrt{-g}\mathcal{L}_{m})}{\delta g_{\mu\nu}}\,. (5)

Since equation (3) must apply for any variation δgμν\delta g_{\mu\nu}, this implies that the tensor field equation becomes

𝒢μν=κTμν.\mathcal{G}^{\mu\nu}=\kappa\,T^{\mu\nu}\,. (6)

In our notation, the μ\nabla_{\mu} symbol is understood as the covariant derivative. Additionally, unless otherwise specified, we use the symbols R/R\partial_{R}\equiv\partial/\partial R and A/A\partial_{A}\equiv\partial/\partial A to denote derivative with respect to RR or AA, respectively.

We recall that many functional forms for g\mathcal{L}_{g} have been proposed to obtain cosmic solutions within the Ricci-inverse theory Amendola et al. (2020); Das et al. (2022).

However, as we are going to show in the next Section III, such models generally fail to explain the late-time acceleration phase of the Universe since their trajectories encounter no-go singularities.

III Ricci-inverse FLRW cosmologies

On sufficiently large scales, the Cosmological Principle states that the universe is uniformly isotropic and homogeneous. The most general space-time compatible with such assumptions is parametrized in terms of the FLRW line-element, which in polar coordinates is given by

ds2=c2dt2+a2(dr2+r2dΩ22),ds^{2}=-c^{2}dt^{2}+a^{2}\!\left(dr^{2}+r^{2}d\Omega_{2}^{2}\right)\,, (7)

where a(t)a(t) is the scale factor and dΩ22=dθ2+sin2θdφ2d\Omega_{2}^{2}=d\theta^{2}+\sin^{2}\!\theta\,d\varphi^{2} is the tridimensional solid-angle element.

We point out that in (7) we assume the flat-space case. In fact, being our study focused on DE (and inflationary) models, observations show that in these casuistics the space-curvature density parameter Ωk\Omega_{k} is very close to zero Aghanim et al. (2020); Hergt et al. (2022). Therefore the space-curvature interferes little or nothing on the actual realization of the no-go theorems Amendola et al. (2020); Do (2022).

Now, the expression for the Ricci scalar calculated from the metric (7) is given by

R=6H2c2(ξ+2),R=\frac{6H^{2}}{c^{2}}(\xi+2)\,, (8)

where the Hubble parameter Ha˙/aH\equiv\dot{a}/a quantifies the rate of time-evolution of the scale factor, and the overdot denotes derivative with respect to cosmic time tt.

For convenience, we have also introduced the new variable ξH/H=a¨a/a˙21\xi\equiv H^{\prime}/H=\ddot{a}\,a/\dot{a}^{2}-1, where the prime symbol stands for =d/dlna{}^{\prime}=d/d\ln a.

According to these results, using definition (1), the corresponding anti-curvature scalar has expression

A=2c23H26+5ξ(1+ξ)(3+ξ).A=\frac{2c^{2}}{3H^{2}}\frac{6+5\xi}{(1+\xi)(3+\xi)}\,. (9)

We end up this section by pointing out: (i) that our results (8) and (9) confirm once again that AR1A\neq R^{-1}, and (ii) that in the Minkowski limit H0H\!\rightarrow\!0 the anti-curvature scalar AA becomes singular while RR and A1A^{-1} do not.

III.1 The first no-go theorem

Before proceeding, it is worth briefly discussing a pathology that afflicts FLRW Ricci-inverse cosmologies.

According to expressions (8) and (9), we see that both the Ricci and anti-curvature scalars exhibit some critical points. In fact, except for special initial conditions such that HH vanishes or diverges: (i) AA is singular for ξc,1=1\xi_{\mbox{\tiny c,1}}=-1 and ξc,3=3\xi_{\mbox{\tiny c,3}}=-3, (ii) RR vanishes for ξc,2=2\xi_{\mbox{\tiny c,2}}=-2, and (iii) AA vanishes for ξc,4=6/5\xi_{\mbox{\tiny c,4}}=-6/5. Due to this fact, it is clear that any Lagrangian density g(R,A)\mathcal{L}_{g}(R,A) containing terms proportional to AA^{\ell} or R||R^{-|\ell|} (with any positive or negative \ell) cannot in general join without problems the cosmic epochs passing through ξc,i\xi_{\mbox{\tiny c,i}} (i=1,2,3,4i=1,2,3,4). Indeed, these terms blow up due to the singularities sourced by ξc,i\xi_{\mbox{\tiny c,i}}, and this fact has repercussions on the equations of motion and their solutions, which will show singularities at the same cosmic epochs Amendola et al. (2020). As we are going to discuss, this fact emerges to be particularly detrimental during the late-time accelerated expansion of universe.

It can be proven that the ξ\xi variabile can be rewritten

ξ=32(1+weff),\xi=-\tfrac{3}{2}\big{(}1+w_{\mbox{\tiny{eff}}}\big{)}\,, (10)

where weffw_{\mbox{\tiny{eff}}} is the effective Equation of State (EoS) parameter of the universe Amendola et al. (2020). Recent observations confirm that our universe has evolved from a decelerated phase weff=0w_{\mbox{\tiny{eff}}}=0 (ξ1.5\xi\approx-1.5) to the current accelerated phase weff=0.685w_{\mbox{\tiny{eff}}}=-0.685 (ξ0.472\xi\approx-0.472) Scolnic et al. (2018b); Amendola et al. (2020). Thus, being ξc,1,ξc,4[1.5;0.472]\xi_{\mbox{\tiny c,1}},\xi_{\mbox{\tiny c,4}}\in[-1.5;-0.472], it is clear that any power of AA evolves a singularity between these two epochs. Due to the above restrictions, no power of RR can cure such singularities and this fact results in a no-go theorem rejecting the Ricci-inverse gravity to be a suitable DE candidate to smoothly joint these two epochs.

Evidence of this fact is well illustrated in Fig. 2, where we plot the cosmic solutions for the Ricci-inverse theory g=R4A1\mathcal{L}_{g}=R-4A^{-1}. Here, the expected singularity sourced by the A1A^{-1} term is ξc,4=1.2\xi_{\mbox{\tiny c,4}}=-1.2, which clearly appears as a divide of the solutions. Consequently, due to the ξc,4\xi_{\mbox{\tiny c,4}} ridge, the cosmic evolution cannot smoothly pass from ξ1.5\xi\approx-1.5 to ξ0.472\xi\approx-0.472.

To author’s knowledge, no valid proposal has been currently found to simply circumvent the action of the first no-go theorem Amendola et al. (2020); Das et al. (2022). However, in searching for a possible solution to the problem, the author realized the existence of a new property that could regularize no-go singularities.

IV A novel dark energy model with complexifiable singularities

In this Section we introduce the notion of regularizability with complexifiable singularities [𝒫2][\mathcal{P}2] and show explicitly how it allows us to get rid of the no-singularities. In mathematical terms, this property refers to the characteristic of certain non-polynomial functions to be able to shift, under certain conditions on their free parameters, their (real) singularities from the real axis \mathbb{R} to the complex plane \mathbb{C}.

Refer to caption
Figure 1: Schematic representation of the circumvention method of the first no-go theorem using complexifiable singularities. If Immξc,i0\mbox{Imm}\,\xi^{\prime}_{\mbox{\tiny c,i}}\neq 0, then the first no-go theorem becomes inoffensive since the cosmic trajectories are real.

With this idea in mind, we can argue as follows. In the case of Ricci-inverse gravity we observe that ξc,i\xi_{\mbox{\tiny c,i}}\in\mathbb{R}, i=1,2,3,4\forall\,i=1,2,3,4. Hence, if we able to designing a Lagrangian density g\mathcal{L}_{g} that satisfies [𝒫2][\mathcal{P}2], then we can transform any troublesome singularity ξc,iξc,i=Reξc,i+jImmξc,i\xi_{\mbox{\tiny c,i}}\in\mathbb{R}\rightarrow\xi^{\prime}_{\mbox{\tiny c,i}}=\mbox{Re}\xi^{\prime}_{\mbox{\tiny c,i}}+j\mbox{Imm}\xi^{\prime}_{\mbox{\tiny c,i}}\in\mathbb{C}. As a consequence, if Immξc,i0\mbox{Imm}\xi^{\prime}_{\mbox{\tiny c,i}}\neq 0 (see, e.g. Fig. 1), then the effects of the first no-go theorem become inoffensive, since cosmic trajectories are physical, i.e. they only evolve along the real axis. In particular, by virtue of this fact, the undesirable singularities that lie between the decelerated and accelerated phases of our universe can be eliminated, resurrecting the Ricci-inverse theory as a suitable DE alternative.

Essential to this scenario is the observation that the design of Lagrangian densities satisfying the [𝒫2][\mathcal{P}2] property is by no means straightforward. For example, the classes of (ruled-out) theories g=RαA1\mathcal{L}_{g}=R-\alpha A^{-1} (see, e.g. Refs. Amendola et al. (2020); Das et al. (2022)) or g=RαA\mathcal{L}_{g}=R-\alpha A (see, e.g. de Souza et al. (2024)), with α\alpha\in\mathbb{R}, do not satisfy [𝒫2][\mathcal{P}2]. Indeed, there is no value of α\alpha able to influence the position of the ξc,1=1\xi_{\mbox{\tiny c,1}}=-1 and ξc,4=6/5\xi_{\mbox{\tiny c,4}}=-6/5 singularities. This fact is easily visible by substituting (8) and (9) in the above Lagrangian densities and then making the common denominator.

IV.1 Circumventing the first no-go theorem

A simple (and novel) model illustrating how our method leads to singularity-free cosmologies considers a Lagrangian density of the form

g=R[1+αβ+AR],\mathcal{L}_{g}=R\left[1+\frac{\alpha}{\beta+AR}\right]\,, (11)

where α,β\alpha,\beta\in\mathbb{R} are constant parameters. It is immediate to check that if α=αGR=0\alpha=\alpha_{GR}=0, then (11) reduces to the usual EH action for GR. Besides, when α0\alpha\neq 0 and β=0\beta=0 we fall back in the (ruled-out) models within the Class Ia studied in Refs. Amendola et al. (2020); Das et al. (2022). In addition, we stress that α\alpha and β\beta are dimensionless. Therefore, our MG model does not require any new dimensional scales, just as it does not introduce any additional fields to those of GR theory.

The strength of our method lies in the fact that, provided 𝒫2\mathcal{P}2 is satisfied, it allows us to very easily transform no-go singularities from real values to complex values, thus rendering them physically ineffective. Now we show that this is exactly what happens. We begin by writing down our Lagrangian (11) in a FLRW universe. Recalling (8) and (9), we easily find that

g=6H2c2(ξ+2)[1+α(ξ+1)(ξ+3)(β+20)ξ2+4(β+16)(4ξ+3)].\mathcal{L}_{g}\!=\!\frac{6H^{2}}{c^{2}}(\xi\!+\!2)\left[1\!+\!\frac{\alpha(\xi\!+\!1)(\xi\!+\!3)}{(\beta\!+\!20)\xi^{2}\!+\!4(\beta\!+\!16)(4\xi\!+\!3)}\right]. (12)

The no-go singularities are clearly evident in the denominator of (12) and solve the algebraic equation of the second degree

(β+20)ξc2+4(β+16)(4ξc+3)=0,(\beta+20)\xi_{c}^{2}+4(\beta+16)(4\xi_{c}+3)=0\,, (13)

whose discriminant is

Δ=4(β+4)(β+16).\Delta=4(\beta+4)(\beta+16)\,. (14)

We now demand that the theory is non-pathological, in the sense that equation (13) admits only complex solutions. To meet this demand, we impose Δ<0\Delta<0, which finally results in the singularity-free condition

16<β<4.-16<\beta<-4\,. (15)

From this result, it can be stated that if β\beta satisfies condition (15), then the FLRW cosmologies emerging form our Lagrangian (11) are safe from the effects of no-go singularities. Interestingly, this condition does not depend on the value of the coefficient α\alpha.

As a final remark, we point out that if, instead of β+AR\beta+AR, in the denominator of (11) we had chosen a higher-degree polynomial like p[βp+(AR)q]p\sum_{p}[\beta_{p}+(AR)^{q}]^{p}, p,qp,q\in\mathbb{R}, then the analogous of equation (13) would become more difficult to discuss being of degree greater than two. Thus, for the illustrative purpose of our study, we decide to adopt the simplest choice (11).

V Cosmological equations

The preceding Subsection IV.1 was dedicated to introduce a novel Lagrangian density (11) that in principle, by imposing condition (15), we expect to regulate the no-go singularities plaguing Ricci-inverse cosmologies. Now we have knowledge of the g\mathcal{L}_{g} profile, we want to find the evolution of the scale factor and other quantities to verify that this is exactly what happens.

To achieve this, we start with the energy-momentum tensor (5), which we assume to be that of a perfect fluid. According to the literature, it can be written as

Tμν=(ρ+c2P)uμuν+Pgμν,T^{\mu\nu}=\left(\rho+c^{-2}P\right)u^{\mu}u^{\nu}+Pg^{\mu\nu}\,, (16)

where ρ=iρ(i)\rho=\sum_{i}\rho_{(i)} and P=iP(i)P=\sum_{i}P_{(i)} are the total mass-energy density and the total hydrostatic pressure of a mixture of two or more non-interacting perfect fluids, each labelled with the index iIi\in I. It is well known that the set II can include a variety of fluids, such as: radiation (r)(r) for photons, pressureless non-relativistic matter (m)(m) for baryonic matter and dark matter, and the cosmological constant (Λ)(\Lambda) for non-dynamical dark energy.

The mass-energy densities and the hydrostatic pressures that appear in Eq. (16) are all measured in the fluid’s (comoving) rest frame, where the fluid’s four-velocity uμu^{\mu} has expression uμ=(c,0,0,0)u^{\mu}=\!^{\top}\!(c,0,0,0). Hence, we can then rewrite the above expression as

Tνμ=diag(ρc2,P,P,P).T^{\mu}_{\nu}=\mbox{diag}\left(-\rho c^{2},P,P,P\right)\,. (17)

In this context, it is useful recalling that the relationship between energy density and pressure is generally described by means of an Equation of State (EoS). Accordingly, henceforth we use assume the following P(ρ)P(\rho) barotropic EoS

P(i)=c2w(i)ρ(i)(iI),P_{(i)}=c^{2}w_{(i)}\rho_{(i)}\quad(i\in I)\,, (18)

which applies for each ii-component of the fluid. Here, w(i)w_{(i)}\in\mathbb{R} are constant parameters. In particular, w(r)=1/3w_{(r)}=1/3 for a radiation density, w(m)=0w_{(m)}=0 for a pressureless non-relativistic matter density, and w(Λ)=1w_{(\Lambda)}=-1 for a cosmological constant density.

Having introduced the necessary background, we can now proceed to derive the modified Friedmann equations.

Using the profiles (7) (11) and expressions (17) (18) and then applying them to the field equation (6), the following cosmological equations for the density parameter Ω(i)83πGρ(i)H2\Omega_{(i)}\equiv\tfrac{8}{3}\pi G\rho_{(i)}H^{-2} can be obtained after somewhat tedious algebra

iIΩ(i)=ψ0+ψ1ξ,\displaystyle\sum_{i\in I}\Omega_{(i)}=\psi_{0}+\psi_{1}\xi^{\prime}\,, (19a)
iIw(i)Ω(i)=φ0+φ1ξ+φ2ξ′′+φ3(ξ)2,\displaystyle\sum_{i\in I}w_{(i)}\Omega_{(i)}=\varphi_{0}+\varphi_{1}\xi^{\prime}+\varphi_{2}\xi^{\prime\prime}+\varphi_{3}(\xi^{\prime})^{2}\,, (19b)
Ω(i)=χ(i)Ω(i)(iI).\displaystyle\Omega_{(i)}^{\prime}=\chi_{(i)}\Omega_{(i)}\quad(i\in I)\,. (19c)

Strictly speaking, Eqs. (19a) and (19b) are derived directly from the (t,t)(t,t) and the (r,r)(r,r) components of the field equation (6), respectively. Using these, we then also derived the matter conservation equation (19c), which emerges because the left-hand side of Eq. (6) is divergenceless, i.e. it satisfies the Bianchi identity, when equipped with our profile (11). However, only two of the three Eqs. (19) are functionally independent.

Interestingly, since Eq. (19c) depends at most on ξ′′\xi^{\prime\prime} and being ξa¨\xi\sim\ddot{a}, our model (as the whole Ricci-inverse framework) is a type of fourth-order gravity. Our system (19) depends on a set of seven ξ\xi-dependent functions, which have the following expression

ψ0(ξ)\displaystyle\psi_{0}(\xi) =1+α(ξ+3)2(β(ξ+1)2+4(ξ+2)2)(β(ξ+1)(ξ+3)+4(ξ+2)(5ξ+6))2,\displaystyle=1+\frac{{\alpha(\xi+3)^{2}(\beta(\xi+1)^{2}+4(\xi+2)^{2})}}{{(\beta(\xi+1)(\xi+3)+4(\xi+2)(5\xi+6))^{2}}}\,, (20)
ψ1(ξ)\displaystyle\psi_{1}(\xi) =(8α(β(ξ(ξ(5ξ+18)+27)+18)+36(ξ+2)3)β(ξ+1)(ξ+3)+4(ξ+2)(5ξ+6)3,\displaystyle=-\frac{(8\alpha(\beta(\xi(\xi(5\xi+18)+27)+18)+36(\xi+2)^{3})}{\beta(\xi+1)(\xi+3)+4(\xi+2)(5\xi+6)^{3}}\,, (21)

and

φ0(ξ)\displaystyle\varphi_{0}(\xi) =(2ξ+3)(α(ξ+3)2(β(ξ+1)2+4(ξ+2)2)+(β(ξ+1)(ξ+3)+4(ξ+2)(5ξ+6)2)3(β(ξ+1)(ξ+3)+4(ξ+2)(5ξ+6))2,\displaystyle=-\frac{{(2\xi+3)(\alpha(\xi+3)^{2}(\beta(\xi+1)^{2}+4(\xi+2)^{2})+(\beta(\xi+1)(\xi+3)+4(\xi+2)(5\xi+6)^{2})}}{{3(\beta(\xi+1)(\xi+3)+4(\xi+2)(5\xi+6))^{2}}}\,, (22)
φ1(ξ)\displaystyle\varphi_{1}(\xi) =8α(ξ+2)(β(ξ(ξ(5ξ+18)+27)+18)+36(ξ+2)3)(β(ξ+1)(ξ+3)+4(ξ+2)(5ξ+6))3,\displaystyle=8\alpha\frac{(\xi+2)(\beta(\xi(\xi(5\xi+18)+27)+18)+36(\xi+2)^{3})}{(\beta(\xi+1)(\xi+3)+4(\xi+2)(5\xi+6))^{3}}\,, (23)
φ2(ξ)\displaystyle\varphi_{2}(\xi) =8α(β(ξ(ξ(5ξ+18)+27)+18)+36(ξ+2)3)3(β(ξ+1)(ξ+3)+4(ξ+2)(5ξ+6))3,\displaystyle=8\alpha\frac{(\beta(\xi(\xi(5\xi+18)+27)+18)+36(\xi+2)^{3})}{3(\beta(\xi+1)(\xi+3)+4(\xi+2)(5\xi+6))^{3}}\,, (24)
φ3(ξ)\displaystyle\varphi_{3}(\xi) =8α(β2(ξ(ξ(ξ(5ξ+24)+54)+72)+45)8β(ξ(ξ+3)(ξ(17ξ+45)+108)+180)720(ξ+2)4β(ξ+1)(ξ+3)+4(ξ+2)(5ξ+6)4,\displaystyle=8\alpha\frac{{(-\beta^{2}(\xi(\xi(\xi(5\xi+24)+54)+72)+45)-8\beta(\xi(\xi+3)(\xi(17\xi+45)+108)+180)-720(\xi+2)^{4}}}{{\beta(\xi+1)(\xi+3)+4(\xi+2)(5\xi+6)^{4}}}\,, (25)

and

χ(i)(ξ)=[3(w(i)+1)+2ξ](iI).\chi_{(i)}(\xi)=-\big{[}3\left(w_{(i)}+1\right)+2\xi\big{]}\quad(i\in I)\,. (26)

Finally, by defining the anti-curvature density parameter Ω(A)=1ψ0+ψ1ξ\Omega_{(A)}=1-\psi_{0}+\psi_{1}\xi^{\prime}, then Eq. (19a) can be rewritten as

iIΩ(i)+Ω(A)=1.\sum_{i\in I}\Omega_{(i)}+\Omega_{(A)}=1\,. (27)

Now we are ready to examine how various cosmological scenarios can be realized within our gravitation model. This topic will be addressed in the next section.

V.1 Power-law & de Sitter solutions

In this section we consider scenarios in which the scale factor varies like (i) a power of the cosmic time, i.e. power-law solutions, or (ii) an exponential of the cosmic time, i.e. a de Sitter (dS) expansion:

a(t){tnwithn/0(Power-law)eH¯twithH¯(de Sitter).a(t)\sim\begin{cases}\,t^{n}\,\,\,\,\,\,\,\mbox{with}\,\,\,\,n\in\mathbb{R}_{/0}&\mbox{(Power-law)}\\ \,e^{\bar{H}t}\,\,\,\,\mbox{with}\,\,\,\,\bar{H}\in\mathbb{R}&\mbox{(de Sitter)}\end{cases}\,. (28)

Here, H¯\bar{H} is the Hubble constant. The present value for the scale factor is assumed to be a0=1a_{0}=1. Both the cases are characterized by ξξ¯=const\xi\equiv\bar{\xi}=const. In particular, we have ξ¯=1/n\bar{\xi}=-1/n for power-law solutions, and ξ¯=0\bar{\xi}=0 for the dS case. Thus, by inversion of Eq. (10), we get the (constant) EoS parameter of the effective fluid to be

w¯eff={(2/3ξ¯+1)(Power-law)1(de Sitter).\bar{w}_{\mbox{\tiny{eff}}}=\begin{cases}-(2/3\,\bar{\xi}+1)\quad&\mbox{(Power-law)}\\ -1\quad&\mbox{(de Sitter)}\end{cases}\,. (29)

We recall that the observed present accelerated value for w¯eff\bar{w}_{\mbox{\tiny{eff}}} is w¯eff=ΩΛ0.685\bar{w}_{\mbox{\tiny{eff}}}=-\Omega_{\Lambda}\approx-0.685 Aghanim et al. (2020). Instead, an EoS parameter w¯eff=0\bar{w}_{\mbox{\tiny{eff}}}=0 corresponds to an effective matter-dominated era.

V.2 Pressureless matter (m)(m)

For the illustrative purpose of our study, we decide to assume the simplest case of an energy-momentum tensor dominated by pressureless matter. This type of fluid, which corresponds to an EoS parameter w(m)=0w_{(m)}=0, represents baryonic matter and cold dark matter.

In this scenario, let us construct the vector x=(ξ,Ω(m))2x=\big{(}\xi,\Omega_{(m)}\big{)}\in\mathbb{R}^{2}. It collects the two dynamical variables of our system (19). In particular, using perturbation theory, we can establish up to the first-order the following decomposition

xx¯+δx,x\sim\bar{x}+\delta x\,, (30)

where

x¯(ξ¯,Ω¯(m))δx(δξ,δΩ(m)),\bar{x}\equiv\!^{\top}\!\big{(}\bar{\xi},\bar{\Omega}_{(m)}\big{)}\qquad\delta x\equiv\!^{\top}\!\big{(}\delta\xi,\delta\Omega_{(m)}\big{)}\,, (31)

with |δξ|1|\delta\xi|\ll 1 and |δΩ(m)|1|\delta\Omega_{(m)}|\ll 1 small perturbations. Clearly, this also shows that any arbitrary function g(x)g(x) can be expanded as gg¯+xg¯δxg\sim\bar{g}+\overline{\nabla\!_{x}g}\cdot\delta x, where the dot denotes cross product. For the sake of convenience of notation, we will assume hereafter that the crossed-out quantities mean evaluation on x¯\bar{x}. Hence, xg¯=xg|x¯\overline{\nabla\!_{x}g}=\nabla\!_{x}g|_{\bar{x}}.

Let us now turn to the modified Friedmann equations with pressureless matter. As a first step, the expansion of equations (19a) and (19b) with the ansatz (30) leads (i) to the background (zeroth-order) system

Ω¯(m)\displaystyle\,\bar{\Omega}_{(m)} =ψ¯0,\displaystyle=\bar{\psi}_{0}\,, (32a)
Ω¯(m)\displaystyle\,\bar{\Omega}^{\prime}_{(m)} =χ¯(m)Ω¯(m),\displaystyle=\bar{\chi}_{(m)}\bar{\Omega}_{(m)}\,, (32b)

and (ii) to the first-order matrix equation

δx=Π¯δx,\delta x^{\prime}=\bar{\Pi}\,\delta x\,, (33)

where

Π¯(ψ¯11ξψ0¯ψ¯11Ω¯(m)ξχ(m)¯χ¯(m)).\bar{\Pi}\equiv\left(\begin{array}[]{cc}-\bar{\psi}_{1}^{-1}\overline{\partial_{\xi}\psi_{0}}&\bar{\psi}_{1}^{-1}\\[2.84544pt] \bar{\Omega}_{(m)}\overline{\partial_{\xi}\chi_{(m)}}&\bar{\chi}_{(m)}\end{array}\right)\,. (34)

In particular, the zero-order system (32) represents the underlying Friedmann equations that govern the evolution of the universe. In contrast, the first-order equation (33) will be useful for performing the linear stability analysis of the background solutions that we will study

V.2.1 Background equations

Let us discuss about solutions of the background system (32). Since ψ¯0\bar{\psi}_{0} is a constant function, then Eq. (35a) dictates that Ω¯(m)\bar{\Omega}_{(m)} is a constant function too. Therefore, we can rewrite the above system as

Ω¯(m)=ψ¯0,\displaystyle\bar{\Omega}_{(m)}=\bar{\psi}_{0}\,, (35a)
χ¯(m)Ω¯(m)=0.\displaystyle\bar{\chi}_{(m)}\bar{\Omega}_{(m)}=0\,. (35b)

We begin our discussion with equation (35b), identifying two main classes of background solutions to study.

The first class (I) is characterized by the condition χ¯(m)=2ξ¯3=0\bar{\chi}_{(m)}=2\bar{\xi}-3=0, leading to the solution (29)

x¯IDM=(32,1+α4+β)w¯effI,DM=0,\bar{x}_{\mbox{\tiny I}}^{\mbox{\tiny DM}}=\!^{\top}\!\left(-\frac{3}{2},1+\frac{\alpha}{4+\beta}\right)\qquad\bar{w}_{\mbox{\tiny{eff}}}^{\mbox{\tiny I,DM}}=0\,, (36)

with β4\beta\neq-4. Interestingly, this condition is already fulfilled as our system is devoid of singularities by condition (15). This solution corresponds to an exact matter-dominated era in which AA acts as a form of dark matter (DM) in addition to a pressureless matter (dust) contribution, whose density parameter is Ω¯(m)I=ψ¯0(3/2)\bar{\Omega}_{(m)}^{\mbox{\tiny I}}=\bar{\psi}_{0}(-3/2). Of course, one can also easily see that if β=0\beta=0, then the well-know result of Amendola et al. obtained in Ref. Amendola et al. (2020) for g=RαA1\mathcal{L}_{g}=R-\alpha A^{-1} is recovered.

Alternatively, the second case (II) that solves Eq. (35) is characterized by the condition Ω¯(m)II=0\bar{\Omega}_{(m)}^{\mbox{\tiny II}}=0, i.e a null pressureless matter contribution to universe’s energy density. In particular, such case reduces to the modified FLRW solution

x¯II=(ξ¯II,0)w¯effII=123ξ¯II,\bar{x}_{\mbox{\tiny II}}=\!^{\top}\!\left(\bar{\xi}_{\mbox{\tiny II}},0\right)\qquad\bar{w}_{\mbox{\tiny{eff}}}^{\mbox{\tiny II}}=-1-\frac{2}{3}\bar{\xi}_{\mbox{\tiny II}}\,, (37)

where ξ¯II\bar{\xi}_{\mbox{\tiny II}} are the real solutions of the algebraic 4-degree equation ψ¯0(ξ¯II)=0\bar{\psi}_{0}(\bar{\xi}_{\mbox{\tiny II}})=0, whose explicit expression is

0\displaystyle 0 =[9α(β+16)+β(9β+320)+2560]\displaystyle=\big{[}9\alpha(\beta+16)+\beta(9\beta+320)+2560\big{]}
+24[(β(α+β)+10α)]ξ¯II\displaystyle+24\big{[}(\beta(\alpha+\beta)+10\alpha)\big{]}\bar{\xi}_{\mbox{\tiny II}}
+2[β(11(α+β)+364)+2(37α+1054)]ξ¯II2\displaystyle+2\big{[}\beta(11(\alpha+\beta)+364)+2(37\alpha+1054)\big{]}\bar{\xi}^{2}_{\mbox{\tiny II}} (38)
+8[α(β+5)+β(β+36)+320]ξ¯II3\displaystyle+8\big{[}\alpha(\beta+5)+\beta(\beta+36)+320\big{]}\bar{\xi}^{3}_{\mbox{\tiny II}}
+[α(β+4)+(β+20)2]ξ¯II4.\displaystyle+\big{[}\alpha(\beta+4)+(\beta+20)^{2}\big{]}\bar{\xi}^{4}_{\mbox{\tiny II}}\,.

Of course, directly addressing the ξ¯II\bar{\xi}_{\mbox{\tiny II}}-solutions of Eq. (V.2.1) is not so simple. Therefore, instead of discussing the (very long) explicit solutions of such equation on varying α,β\alpha,\beta parameters, we proceed (i) by first selecting the α,β\alpha,\beta values that impose two ξ¯II\bar{\xi}_{\mbox{\tiny II}}-solutions of Eq. (V.2.1) to be the cosmologies we wish to reproduce, and (ii) then directly solving the residual algebraic 2nd-degree equation.

First of all, to lower the degree of the equation Eq. (V.2.1), we decide to force one solution to be a dS state. On the other hand, let’s remember that the second no-go theorem for Ricci-inverse gravity considers the dS case, which then emerges as an intriguing solution. Moreover, as we shall see in Subsection V.2.2, this choice will turn out to be particularly apt, being such cosmology a stable attractor of our model.

So, following the prescription outlined before, one can straightforwardly decide that a first solution could be

x¯IIdS=(0,0)w¯effII,dS=1.\bar{x}^{\mbox{\tiny dS}}_{\mbox{\tiny II}}=\!^{\top}\!\left(0,0\right)\qquad\bar{w}_{\mbox{\tiny{eff}}}^{\mbox{\tiny II,dS}}=-1\,. (39)

Moving to a second cosmological solution, we can’t help but think about the observed present value w¯eff=ΩΛ0.685\bar{w}_{\mbox{\tiny{eff}}}=-\Omega_{\Lambda}\approx-0.685 Aghanim et al. (2020). Hence, by using Eq. (29), such scenario corresponds to ξ¯IIDE0.473\bar{\xi}^{\mbox{\tiny DE}}_{\mbox{\tiny II}}\approx-0.473, which in terms of x¯\bar{x} reduces to the following second solution

x¯IIDE=(0.473,0)w¯effII,DE=0.685.\bar{x}^{\mbox{\tiny DE}}_{\mbox{\tiny II}}=\!^{\top}\!\left(-0.473,0\right)\qquad\bar{w}_{\mbox{\tiny{eff}}}^{\mbox{\tiny II,DE}}=-0.685\,. (40)

Using solutions (39),(40) and subsequently computing them into the polynomial equation (V.2.1), the following values for α\alpha and β\beta can be obtained

α0.028β15.972.\alpha\approx-0.028\qquad\beta\approx-15.972\,. (41)

Interestingly, the value found for α\alpha is very close to zero. As far as our g\mathcal{L}_{g} model is concerned, this fact results in a very week coupling of the anti-curvature contribute (β+AR)1(\beta+AR)^{-1} to the gravitational parameter GG. However, although small, this value contributes significantly to the reproduction of the present accelerated expansion of universe. Moreover, the value of β\beta - to our delight - satisfies the condition (15) that makes our model singularity-free with respect to the cosmological solutions x¯IIdS\bar{x}^{\mbox{\tiny dS}}_{\mbox{\tiny II}} and x¯IIDE\bar{x}^{\mbox{\tiny DE}}_{\mbox{\tiny II}}.

With these values in mind, Eq. (V.2.1) becomes

ξ(ξξ¯IIDE)(ξ2+ε1ξ+ε0)=0,\xi\left(\xi-\bar{\xi}^{\mbox{\tiny DE}}_{\mbox{\tiny II}}\right)\left(\xi^{2}+\varepsilon_{1}\xi+\varepsilon_{0}\right)=0\,, (42)

with ε10.271\varepsilon_{1}\approx-0.271 and ε10.514\varepsilon_{1}\approx 0.514. The solutions of the residual 2nd-degree polynomial equation appearing in Eq. (42) are ξ¯II±0.135±0.704j\bar{\xi}^{\pm}_{\mbox{\tiny II}}\approx 0.135\pm 0.704j. It is clear that such solutions are complex numbers, so they are not physical. We recall that jj is the imaginary unit.

We can now study qualitatively our solutions by determining their stability.

V.2.2 Linear stability analysis

In this subsection we will use the linear stability theory to investigate the stability of our cosmological solutions x¯IDM\bar{x}^{\mbox{\tiny DM}}_{\mbox{\tiny I}}, x¯IIdS\bar{x}^{\mbox{\tiny dS}}_{\mbox{\tiny II}} and x¯IIDE\bar{x}^{\mbox{\tiny DE}}_{\mbox{\tiny II}}, which may been seen as critical points for the dynamical system (19). This method studies the eigenvalues λ¯\bar{\lambda} of the matrix Π¯\bar{\Pi} defined in (34) and, in particular, analyzes the sign of their real part. On general grounds, if Reλ¯<0\mbox{Re}\bar{\lambda}<0, λ¯\forall\bar{\lambda}, then x¯\bar{x} is considered a stable attractor solution of the dynamical system.

By definition, the eigenvalues associated to Π¯\bar{\Pi} can be easily determined by finding the roots of the characteristic polynomial

pΠ¯(x¯,λ¯)\displaystyle p_{\bar{\Pi}}(\bar{x},\bar{\lambda}) det(Π¯λ¯𝕀2)\displaystyle\equiv\det\left(\bar{\Pi}-\bar{\lambda}\mathbb{I}_{2}\right) (43)
=(ψ¯11ξψ0¯+λ¯)(χ¯(m)λ¯)+ψ¯11Ω¯(m)ξχ(m)¯.\displaystyle=\left(\bar{\psi}_{1}^{-1}\overline{\partial_{\xi}\psi_{0}}+\bar{\lambda}\right)\!\left(\bar{\chi}_{(m)}-\bar{\lambda}\right)+\bar{\psi}_{1}^{-1}\bar{\Omega}_{(m)}\overline{\partial_{\xi}\chi_{(m)}}\,.

As the notation suggests, the characteristic polynomial takes on a different expression depending on the solution x¯\bar{x} considered. So, let us now move onto the specific cases.

We begin from the DM scenario (36). Searching for the eigenvalues solving pΠ(x¯IDM,λ¯IDM)=0p_{\Pi}(\bar{x}_{\mbox{\tiny I}}^{\mbox{\tiny DM}},\bar{\lambda}_{\mbox{\tiny I}}^{\mbox{\tiny DM}})=0, we get

λ¯I,±DM=34±18α6α[α(β2)+(β+4)2].\bar{\lambda}_{\mbox{\tiny I},\pm}^{\mbox{\tiny DM}}=\frac{3}{4}\pm\frac{1}{8\alpha}\sqrt{-6\alpha\big{[}\alpha(\beta-2)+(\beta+4)^{2}\big{]}}\,. (44)

After somewhat tedious algebra, the coexistence of the stability condition λ¯I,±DM<0\bar{\lambda}_{\mbox{\tiny I},\pm}^{\mbox{\tiny DM}}<0 with the singularity-free condition (15) leads to the following constraint

12(α+8+α(α+24))<β<(α+4),-\frac{1}{2}\left(\alpha+8+\sqrt{\alpha(\alpha+24)}\right)<\beta<-(\alpha+4)\,, (45)

with 0<α<80<\alpha<8. In this situation, one can be easily prove that our parameters (41) do not satisfy the constraint (45). This demonstrates that x¯IDM\bar{x}^{\mbox{\tiny DM}}_{\mbox{\tiny I}} is not a stable solution for our model.

In the same way, for the case (II) we find that the roots of pΠ(x¯IIDE,dS,λ¯IIDE,dS)=0p_{\Pi}(\bar{x}_{\mbox{\tiny II}}^{\mbox{\tiny DE,dS}},\bar{\lambda}_{\mbox{\tiny II}}^{\mbox{\tiny DE,dS}})=0 are

λ¯II,1DE,dS=(1+3ξ¯IIDE,dS)λ¯II,2DE,dS=(2+3ξ¯IIDE,dS).\bar{\lambda}_{\mbox{\tiny II},1}^{\mbox{\tiny DE,dS}}=-\big{(}1+3\bar{\xi}^{\mbox{\tiny DE,dS}}_{\mbox{\tiny II}}\big{)}\quad\bar{\lambda}_{\mbox{\tiny II},2}^{\mbox{\tiny DE,dS}}=-\big{(}2+3\bar{\xi}^{\mbox{\tiny DE,dS}}_{\mbox{\tiny II}}\big{)}\,. (46)

Again, after somewhat tedious algebra, we obtain that the overlap between the stability domain λ¯II,1DE,dS<0\bar{\lambda}_{\mbox{\tiny II},1}^{\mbox{\tiny DE,dS}}<0, λ¯II,2DE,dS<0\bar{\lambda}_{\mbox{\tiny II},2}^{\mbox{\tiny DE,dS}}<0 and the singularity-free condition (15) is:

ξ¯IIDE,dS>32,\bar{\xi}^{\mbox{\tiny DE,dS}}_{\mbox{\tiny II}}>-\frac{3}{2}\,, (47)

with 16<β<4-16<\beta<-4. In this situation, one can be easily see that our solutions are fully consistent with the above stability constraint (47). This demonstrates that x¯IIdS\bar{x}^{\mbox{\tiny dS}}_{\mbox{\tiny II}} and x¯IIDE\bar{x}^{\mbox{\tiny DE}}_{\mbox{\tiny II}} are stable attractors of our model.

For a general overview, our findings are summarized in Table I.

Refer to caption
Figure 2: Numerical solutions ξ(a)\xi(a) for Eq. (19b) in the pressureless matter case. Here, α=4\alpha=-4 and β=0\beta=0 do not satisfy the singularity-free condition (15). It is evident that the solutions ξ¯=1.5\bar{\xi}=-1.5 (magenta lines) and ξ¯=0.75\bar{\xi}=-0.75 (blue lines) are stable attractor. However, as discussed in Section III.1, a divide at ξ¯=1.2\bar{\xi}=-1.2 realizes the effects of the no-go theorem.
Refer to caption
Figure 3: Numerical solutions ξ(a)\xi(a) for Eq. (19b) in the pressureless matter case. Here, α=0.028\alpha=-0.028 and β=15.972\beta=-15.972 satisfy both the singularity-free condition (15) and the stability conditions. It is evident that (i) the observed accelerated expansion ξ¯=0.472\bar{\xi}=-0.472 is a stable attractor, as well as (ii) the decelerated phase ξ¯=1.5\bar{\xi}=-1.5 is an unstable solution. As expected, the cosmic trajectories smoothly join these two epochs without falling into no-go singularities.

Graphical evidence that our model works well is shown in Fig. 3, where we solved numerically the evolution equation (19b) for ξ(a)\xi(a). In particular, we observe very good agreement with the required behavior of a cosmic expansion moving from a decelerated phase (around w¯eff=0\bar{w}_{\mbox{\tiny{eff}}}=0) to an accelerated phase (w¯eff=0.685\bar{w}_{\mbox{\tiny{eff}}}=-0.685) without falling into no-go singularities. Moreover, it is evident that the cosmology describing the observed expansion of our universe is a stable attractor.

A further observation can be made. Since x¯IDM\bar{x}^{\mbox{\tiny DM}}_{\mbox{\tiny I}} is not a stable solution, the event that may have initiated cosmic evolution from this era may have been a small perturbation of the coupling constant α\alpha. To prove this, suppose we start from GR theory (recovered from our model with αGR=0\alpha_{\mbox{\tiny GR}}=0), where the matter-dominated era w¯eff=0\bar{w}_{\mbox{\tiny{eff}}}=0 is a stable cosmological solution. On the basis of the parameterization of our model, suppose we now introduce a small perturbation on alphaGRalpha_{\mbox{\tiny GR}} such that αGRαGR+δα\alpha_{\mbox{\tiny GR}}\rightarrow\alpha_{\mbox{\tiny GR}}+\delta\alpha with δα0.028\delta\alpha\approx-0.028. This fact, as seen in the previous discussion, makes the w¯eff=0\bar{w}_{\mbox{\tiny{eff}}}=0 epoch unstable. Therefore, as a result of such a perturbation, our Ricci-inverse (11) is established as a new MG theory in which the anti-curvature AA, acting as a form of DE, smoothly join that phase with the current expansion. Of course, this involves a change in the coupling constant κ\kappa, i.e., the gravitational parameter GG, of the type κ=κGR+δκ(α)\kappa=\kappa_{GR}+\delta\kappa(\alpha).

As a final remark, we point out that if in our model we had considered a different observed value, however close to 0.7-0.7, for w¯eff\bar{w}_{\mbox{\tiny{eff}}}, then the deductions obtainable, apart from the specific values for α\alpha and β\beta, would be the same as in the case discussed in this paper. We also leave to future work any discussion of the existence of existence of ghosts or other types of instabilities related to our model.

DM w¯eff=0\bar{w}_{\mbox{\tiny{eff}}}=0 ξ¯=1.5\bar{\xi}=-1.5 Ω¯(m)=1.002\bar{\Omega}_{(m)}=1.002 Unstable
dS w¯eff=1\bar{w}_{\mbox{\tiny{eff}}}=-1 ξ¯=0\bar{\xi}=0 Ω¯(m)=0\bar{\Omega}_{(m)}=0 Stable
DE w¯eff=0.685\bar{w}_{\mbox{\tiny{eff}}}=-0.685 ξ¯=0.472\bar{\xi}=-0.472 Ω¯(m)=0\bar{\Omega}_{(m)}=0 Stable
Table 1: Constant solutions for Eq. (32) with α0.028\alpha\approx-0.028 and β15.972\beta\approx-15.972. Different epochs of the universe dominated by effective fluids of dark matter (DM), dark energy (DE) and de Sitter (dS) are considered.

VI Conclusions

In the present work, a novel MG theory called Ricci-inverse gravity was investigated. This theory is based on the introduction of a new, purely geometric object called the anti-curvature tensor, defined as the inverse of the well-known Ricci tensor. From this foundation, it is possible to modify the EH action without introducing new fields or requiring a fine-tuning of the coupling constants.

Although many studies emphasize the increasing relevance and applicability of Ricci-inverse gravity in diverse scenarios, some analyses have fed the idea that Ricci-inverse gravity is not a promising theory to describe the current accelerated expansion of the universe. Foremost among them, later translated into terms of a blocking no-go theorem, is the presence of singularities along the path of cosmic trajectories connecting a decelerated (matter-dominated) epoch to the current accelerated expansion of the universe. By requiring the Lagrangian density of such theories to satisfy a new property that we have called regularization via complexification of singularities (𝒫2\mathcal{P}2), we have shown how it is possible to evade the effects of the aforementioned no-go theorem and reconsider cosmologies based on anti-curvature.

In addition to providing a formal introduction to 𝒫2\mathcal{P}2, in this paper we have designed and analyzed for the first time a new scale-free Ricci-inverse gravity model that, by implementing 𝒫2\mathcal{P}2, (i) is not rent by no-go singularities and (ii) brings out the current observed expansion as a stable attractor solution. In particular, after obtaining the modified Friedmann equations, we approached the study of their solutions from both an analytical and numerical point of view, giving evidence that they are regular up to the present observed values.

We trust that, with this analysis, we can once again reopen the discussion aimed at reconsidering gravity theories based on anti-curvature as good dark energy models.

We leave for future investigation the analysis of ghost or possible other instabilities.

Acknowledgements.
The author would like to thank Dr. E. Foltran for computational support during the calculations. Some algebraic computations in this article were performed using the Cadabra software Peeters (2007).

References

  • Riess et al. (1998) A. G. Riess et al. (Supernova Search Team), Astron. J. 116, 1009 (1998), eprint astro-ph/9805201.
  • Abbott et al. (2019) T. M. C. Abbott, S. Allam, P. Andersen, C. Angus, J. Asorey, A. Avelino, S. Avila, B. A. Bassett, K. Bechtol, G. M. Bernstein, et al., The Astrophysical Journal Letters 872, L30 (2019), URL https://dx.doi.org/10.3847/2041-8213/ab04fa.
  • Scolnic et al. (2018a) D. M. Scolnic et al. (Pan-STARRS1), Astrophys. J. 859, 101 (2018a), eprint 1710.00845.
  • Starobinsky (1980) A. A. Starobinsky, Phys. Lett. B 91, 99 (1980).
  • Guth (1981a) A. H. Guth, Phys. Rev. D 23, 347 (1981a), URL https://link.aps.org/doi/10.1103/PhysRevD.23.347.
  • Hinshaw et al. (2013) G. Hinshaw, D. Larson, E. Komatsu, D. N. Spergel, C. L. Bennett, J. Dunkley, M. R. Nolta, M. Halpern, R. S. Hill, N. Odegard, et al., The Astrophysical Journal Supplement Series 208, 19 (2013), ISSN 1538-4365, URL http://dx.doi.org/10.1088/0067-0049/208/2/19.
  • Malik et al. (2023a) A. Malik, E. Meer, Z. Asghar, and A. Ali, Chin. J. Phys. 86, 391 (2023a).
  • Nojiri et al. (2017) S. Nojiri, S. D. Odintsov, and V. K. Oikonomou, Phys. Rept. 692, 1 (2017), eprint 1705.11098.
  • Amendola et al. (2020) L. Amendola, L. Giani, and G. Laverda, Physics Letters B 811, 135923 (2020), ISSN 0370-2693, URL https://www.sciencedirect.com/science/article/pii/S0370269320307267.
  • Shamir et al. (2023) M. F. Shamir, M. Ahmad, G. Mustafa, and A. Rashid, Chin. J. Phys. 81, 51 (2023), eprint 2304.07209.
  • Shamir and Meer (2023) M. F. Shamir and E. Meer, Eur. Phys. J. C 83, 49 (2023), eprint 2304.04194.
  • Malik et al. (2023b) A. Malik, A. Arif, and M. F. Shamir, Int. J. Theor. Phys. 62, 243 (2023b).
  • Malik et al. (2024a) A. Malik, A. Arif, and M. F. Shamir, Eur. Phys. J. Plus 139, 67 (2024a).
  • Malik et al. (2024b) A. Malik, A. Hussain, M. F. Shamir, and A. Almas (2024b), eprint 2404.15203.
  • Jawad and Sultan (2022) A. Jawad and A. M. Sultan, EPL 138, 29001 (2022).
  • Mustafa et al. (2024) G. Mustafa, F. Javed, S. K. Maurya, and A. Errehymy (2024), eprint 2404.04649.
  • Ahmed and Guvendi (2024) F. Ahmed and A. Guvendi, Chin. J. Phys. 89, 69 (2024).
  • de Souza and Santos (2023) J. C. R. de Souza and A. F. Santos, Eur. Phys. J. C 83, 834 (2023), eprint 2309.05439.
  • de Souza et al. (2024) J. C. R. de Souza, A. F. Santos, and F. Ahmed, Eur. Phys. J. C 84, 559 (2024).
  • Scomparin (2021) M. Scomparin (2021), eprint 2102.04676.
  • Aghanim et al. (2020) N. Aghanim et al. (Planck), Astron. Astrophys. 641, A6 (2020), [Erratum: Astron.Astrophys. 652, C4 (2021)], eprint 1807.06209.
  • Das et al. (2022) I. Das, J. P. Johnson, and S. Shankaranarayanan, Eur. Phys. J. Plus 137, 1265 (2022), eprint 2108.00992.
  • Guth (1981b) A. H. Guth, Phys. Rev. D 23, 347 (1981b).
  • Linde (1982) A. D. Linde, Phys. Lett. B 108, 389 (1982).
  • Do (2021) T. Q. Do, Eur. Phys. J. C 81, 431 (2021), eprint 2009.06306.
  • Do (2022) T. Q. Do, Eur. Phys. J. C 82, 15 (2022), eprint 2101.08538.
  • Lee (2021) S. Lee, Journal of Cosmology and Astroparticle Physics 2021, 054 (2021), URL https://dx.doi.org/10.1088/1475-7516/2021/08/054.
  • Hergt et al. (2022) L. T. Hergt, F. J. Agocs, W. J. Handley, M. P. Hobson, and A. N. Lasenby, Phys. Rev. D 106, 063529 (2022), eprint 2205.07374.
  • Scolnic et al. (2018b) D. M. Scolnic et al. (Pan-STARRS1), Astrophys. J. 859, 101 (2018b), eprint 1710.00845.
  • Peeters (2007) K. Peeters, Comput. Phys. Commun. 176, 550 (2007), eprint cs/0608005.