14
Eur. Phys. J. C (2020) 80:286 https://doi.org/10.1140/epjc/s10052-020-7828-7 Regular Article - Theoretical Physics Dynamical systems methods and statender diagnostic of interacting vacuum energy models Grigoris Panotopoulos 1,a , Ángel Rincón 2 ,b , Giovanni Otalora 2 ,c , Nelson Videla 2 ,d 1 Centro de Astrofísica e Gravitação, Instituto Superior Técnico-IST, Universidade de Lisboa-UL, Av. Rovisco Pais, 1049-001 Lisboa, Portugal 2 Instituto de Física, Pontificia Universidad Católica de Valparaíso, Avenida Brasil 2950, Casilla 4059, Valparaiso, Chile Received: 9 November 2019 / Accepted: 11 March 2020 / Published online: 30 March 2020 © The Author(s) 2020 Abstract We study three interacting dark energy models within the framework of four-dimensional General Relativ- ity and a spatially flat Universe. In particular, we first con- sider two vacuum models where dark energy interacts with dark matter, while relativistic matter as well as baryons are treated as non-interacting fluid components. Secondly, we investigate a third model where the gravitational coupling is assumed to be a slowly-varying function of the Hubble rate and dark energy and dark matter interact as well. We com- pute the statefinders parameters versus red-shift as well as the critical points and their nature applying dynamical systems methods. In the case of only an interaction term, our main findings indicate that (i) significant differences between the models are observed as we increase the strength of the inter- action term, and (ii) all the models present an unique attractor corresponding to acceleration. On the other hand, when we allow for a variable gravitational coupling, we find that (i) the deviation from the concordance model depends of both the strength of gravitational coupling parameter and the interac- tion term, and (ii) there is an unique attractor corresponding to acceleration. 1 Introduction The origin and nature of dark energy (DE), the fluid compo- nent that currently accelerates the Universe [13], is one of the biggest mysteries and challenges in modern theoretical Cosmology. Clearly, Einstein’s General Relativity [4] with radiation and matter only cannot lead to accelerating solu- tions. A positive cosmological constant [5] is the simplest, a e-mail: [email protected] (corresponding author) b e-mail: [email protected] c e-mail: [email protected] d e-mail: [email protected] most economical model in a very good agreement with a great deal of current observational data. Since, however, it suffers from the cosmological constant (CC) problem [6], other possibilities have been considered in the literature over the years. The CC problem, introduced by Zeldovich for the first time more than fifty years ago [7], may be summarized in a few words as follows: It is an impressive mismatch-by many orders of magnitude-between the observational value of vac- uum energy, and the expected value from particle physics due to vacuum fluctuations of massive fields. Although some progress has been made up to now, see e.g. [811], the origin of the CC problem still remains a mystery. Regarding the CC problem and possible alternatives to the CDM model, either a modified theory of gravity is assumed, providing correction terms to GR at cosmologi- cal scales, or a new dynamical degree of freedom with an equation-of-state (EOS) parameter w< 1/3 must be intro- duced. In the first class of models (geometrical DE) one finds for instance f ( R) theories of gravity [1215], brane-world models [1618] and Scalar–Tensor theories of gravity [1922], while in the second class (dynamical DE) one finds mod- els such as quintessence [23], phantom [24], quintom [25], tachyonic [26] or k-essence [27]. For an excellent review on the dynamics of dark energy see e.g. [28]. Furthermore, regarding the value of the Hubble constant H 0 , there is nowadays a tension between high red-shift CMB data and low red-shift data, see e.g. [2932]. The value of the Hubble constant extracted by the PLANCK Collabo- ration [33, 34], H 0 = (67 68) km/(Mpc s), is found to be lower than the value obtained by local measurements, H 0 = (73 74) km/(Mpc s) [35, 36]. This tension might call for new physics [37]. What is more, regarding large scale structure formation data, the growth rate from red-shift space distortion measurements has been found to be lower than expected from PLANCK [38, 39]. Both tensions may be alleviated within the framework of running vacuum dynamics [4046]. In this class of models, 123

Dynamical systems methods and statender ... - Springer

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Dynamical systems methods and statender ... - Springer

Eur. Phys. J. C (2020) 80:286https://doi.org/10.1140/epjc/s10052-020-7828-7

Regular Article - Theoretical Physics

Dynamical systems methods and statender diagnosticof interacting vacuum energy models

Grigoris Panotopoulos1,a, Ángel Rincón2,b, Giovanni Otalora2,c, Nelson Videla2,d

1 Centro de Astrofísica e Gravitação, Instituto Superior Técnico-IST, Universidade de Lisboa-UL, Av. Rovisco Pais, 1049-001 Lisboa, Portugal2 Instituto de Física, Pontificia Universidad Católica de Valparaíso, Avenida Brasil 2950, Casilla 4059, Valparaiso, Chile

Received: 9 November 2019 / Accepted: 11 March 2020 / Published online: 30 March 2020© The Author(s) 2020

Abstract We study three interacting dark energy modelswithin the framework of four-dimensional General Relativ-ity and a spatially flat Universe. In particular, we first con-sider two vacuum models where dark energy interacts withdark matter, while relativistic matter as well as baryons aretreated as non-interacting fluid components. Secondly, weinvestigate a third model where the gravitational coupling isassumed to be a slowly-varying function of the Hubble rateand dark energy and dark matter interact as well. We com-pute the statefinders parameters versus red-shift as well as thecritical points and their nature applying dynamical systemsmethods. In the case of only an interaction term, our mainfindings indicate that (i) significant differences between themodels are observed as we increase the strength of the inter-action term, and (ii) all the models present an unique attractorcorresponding to acceleration. On the other hand, when weallow for a variable gravitational coupling, we find that (i) thedeviation from the concordance model depends of both thestrength of gravitational coupling parameter and the interac-tion term, and (ii) there is an unique attractor correspondingto acceleration.

1 Introduction

The origin and nature of dark energy (DE), the fluid compo-nent that currently accelerates the Universe [1–3], is one ofthe biggest mysteries and challenges in modern theoreticalCosmology. Clearly, Einstein’s General Relativity [4] withradiation and matter only cannot lead to accelerating solu-tions. A positive cosmological constant [5] is the simplest,

a e-mail: [email protected] (correspondingauthor)b e-mail: [email protected] e-mail: [email protected] e-mail: [email protected]

most economical model in a very good agreement with agreat deal of current observational data. Since, however, itsuffers from the cosmological constant (CC) problem [6],other possibilities have been considered in the literature overthe years. The CC problem, introduced by Zeldovich for thefirst time more than fifty years ago [7], may be summarized ina few words as follows: It is an impressive mismatch-by manyorders of magnitude-between the observational value of vac-uum energy, and the expected value from particle physicsdue to vacuum fluctuations of massive fields. Although someprogress has been made up to now, see e.g. [8–11], the originof the CC problem still remains a mystery.

Regarding the CC problem and possible alternatives tothe �CDM model, either a modified theory of gravity isassumed, providing correction terms to GR at cosmologi-cal scales, or a new dynamical degree of freedom with anequation-of-state (EOS) parameter w < −1/3 must be intro-duced. In the first class of models (geometrical DE) one findsfor instance f (R) theories of gravity [12–15], brane-worldmodels [16–18] and Scalar–Tensor theories of gravity [19–22], while in the second class (dynamical DE) one finds mod-els such as quintessence [23], phantom [24], quintom [25],tachyonic [26] or k-essence [27]. For an excellent review onthe dynamics of dark energy see e.g. [28].

Furthermore, regarding the value of the Hubble constantH0, there is nowadays a tension between high red-shift CMBdata and low red-shift data, see e.g. [29–32]. The value ofthe Hubble constant extracted by the PLANCK Collabo-ration [33,34], H0 = (67 − 68) km/(Mpc s), is found tobe lower than the value obtained by local measurements,H0 = (73−74) km/(Mpc s) [35,36]. This tension might callfor new physics [37]. What is more, regarding large scalestructure formation data, the growth rate from red-shift spacedistortion measurements has been found to be lower thanexpected from PLANCK [38,39].

Both tensions may be alleviated within the framework ofrunning vacuum dynamics [40–46]. In this class of models,

123

Page 2: Dynamical systems methods and statender ... - Springer

286 Page 2 of 14 Eur. Phys. J. C (2020) 80 :286

contrary to a rigid cosmological constant � = const., vac-uum energy density can be expressed as a function of theHubble rate, i.e. ρ� = ρ�(H), being of dynamical natureand at the same time it may interacts with dark matter, andalso it accounts for a running of the gravitational coupling G[47]. We remark in passing that other alternatives approachesto running vacuum dynamics do exist, and one may mentionfor instance the scale-dependent (SD) scenario [48–50], inwhich it is assumed that the couplings of the original clas-sical action acquire a scale-dependence. The SD scenariois one of the approaches to quantum gravity, inspired bythe well-known Brans–Dicke theory [19,20], where New-ton’s constant is replaced by an dynamical scalar field fol-lowing the identification φ → G−1. In SD cosmologicalmodels the cosmological constant becomes time dependentsimilarly to the running vacuum dynamics, although in theSD scenario Newton’s constant, too, acquires a time depen-dence.

Remarkably, measurements of the expansion rate basedon Hubble-diagram of high-redshift objects [51,52] suggestthat a rigid � term is ruled out by a statistical significanceof ∼ 4σ , accounting for deviations from �CDM model.The aforementioned deviations allow the possibility of bothdynamical and interacting DE, which is realizable within theframework of running vacuum scenario [46]. More generi-cally, interacting DE models are interesting for several rea-sons. First of all, it is a possibility that should not be ignored,under the assumption that DE and DM do not evolve sepa-rately but interact with each other non-gravitationally. Sec-ondly, and perhaps the main motivation for an interactionin the dark sector, is currently motivated that this scenariocan solve the current cosmological tensions in some data,see e.g. [53–55]. Additionally, other recent relevant resultsregarding the interaction between DE and DM were foundin [56,57]. For an extensive review on DE and DM interac-tions, see [58] an references therein. In addition, the “whynow problem” may be addressed if our current Universe sits ata stable fixed point (attractor) of the corresponding dynam-ical system, and this attractor corresponds to accelerationand to 0 < �m,0 < 1, with �m,0 being today’s normal-ized density of matter. Thus, the system will always reachits attractor at late times irrespectively of the initial condi-tions. It can be easily shown that this scenario cannot berealized if there is no interaction between DE and matter[59].

As several DE models predict very similar expansion his-tories, all of them are still in agreement with the availableobservational data. It thus becomes clear that it is advanta-geous to introduce and study new appropriate quantities capa-ble of discriminating between different dark energy cosmo-logical models at least at background level. Hence, in orderto compare different dark energy models we can introduceparameters in which derivatives of the scale factor beyond

the second-order appear. To this end, one option would beto study the so-called statefinder parameters, r, s, defined asfollows [60,61]

r ≡...a

aH3 , (1)

s ≡ r − 1

3(q − 12 )

, (2)

where the dot denotes differentiation with respect to thecosmic time t , H = a/a is the Hubble parameter, andq = −a/(aH2) is the decelerating parameter. We see thatthe statefinder parameters are expressed in terms of the thirdderivative of the scale factor with respect to the cosmic time,contrary to the Hubble parameter and the decelerating param-eter, which are expressed in terms of the first and the sec-ond time derivative of the scale factor, respectively. It isstraightforward to verify that for the �CDM model with-out radiation the statefinder parameters take constant values,r = 1, s = 0. These parameters may be computed withina certain model, their values can be extracted from futureobservations [62,63], and the statefinder diagnostic has beenapplied to several dark energy models [64–68]. As we willsee later on, r, s can be very different from one model toanother even if they predict very similar expansion histories.

Considering that running vacuum (RVM) models offersan interesting framework to study phenomenology beyond to�CDM model, the main goal of the present work is to anal-yse three models within the running vacuum dynamics: wefirst consider two vacuum models where dark energy inter-acts with dark matter [46], and secondly, we investigate athird model where the gravitational coupling is assumed tobe a slowly-varying function of the Hubble rate [47] and darkenergy and dark matter interact as well. The analysis is per-formed in two respects: On the one hand, by applying thedynamical systems methods, we compute the critical pointsfor each scenario and study their stability. On the other hand,in other to discriminate between the several running vacuumDE models and �CDM, we perform the statefinder diag-nostic by means computing the statefinder parameters as afunction of the redshift, studying their high and low-redshiftlimits. Our work is organized as follows: after this introduc-tion, we present the basic equations and analytical solutionsfor Models I and II in Sects. 2 and 3, respectively. In thefourth section, upon the dynamical system analysis we com-pute the corresponding critical points for Models I and II,while in the fifth section we discuss the statefinder param-eters for the same models. In Sect. 6, we present the mainresults for Model III regarding the dynamical system andstatefinder analysis. Finally we summarize our findings andpresent our conclusions in Sect. 7. We adopt the mostly pos-itive metric signature, (−,+,+,+), and we work in naturalunits where c = h = 1.

123

Page 3: Dynamical systems methods and statender ... - Springer

Eur. Phys. J. C (2020) 80 :286 Page 3 of 14 286

2 Theoretical framework

We consider a flat (k = 0) FLRW Universe

ds2 = −dt2 + a(t)2δi j dxi dx j , (3)

and setting κ2 = 8πG, with G being the Newton’s constant,the scale factor a(t) satisfies the Friedmann equations

H2 = κ2

3

A

ρA, (4)

H = −κ2

2

A

(ρA + pA), (5)

where ρA and pA denote the energy density and pressure ofeach individual fluid component, respectively. The equation-of-state parameter for each fluid component pA = wAρA

takes the values: w = 0 for baryons and dark matter, w = 1/3for radiation and w = −1 for DE.

The system of cosmological equations also includesthe conservation equations for the non-interacting fluids(baryons, radiation)

ρb + 3Hρb = 0, (6)

ρr + 4Hρr = 0, (7)

as well as for the interacting components (DE and dark mat-ter)

ρ� = −Q, (8)

ρdm + 3Hρdm = Q. (9)

Here, Q represents the source term, i.e. the energy exchangebetween DE and DM. Particularly, in the running vacuumcosmology scenario, the cosmological coupling � varies as� ≡ �(H2) or � ≡ �(R) [69]. In such a models, thedynamics of vacuum is due to the energy exchange with someof the fluid components that participate to the evolution ofthe Universe. As running vacuum models (RVM) seem toperform better than the �CDM in some circumstances, inthe present work we first consider two scenarios, namely Iand II, found e.g. in [46] and precisely labelled as “runningvacuum model” (RVM):

Q1 = 3νdmHρdm, (10)

Q2 = 3ν�Hρ�, (11)

where the dimensionless parameters {νi }measure the strengthof the interaction term Qi . In the present paper we are inter-ested in studying the late-times cosmology within the inter-acting vacuum energy scenarios. So, we have neglected thecoupling to radiation and baryons because at lower redshifts,z � 2, the contribution to the total energy density comingfrom these components is smaller than the dark energy anddark matter components. It is worth to mention that an even-tual coupling between radiation and dark energy could have a

significant effect on the dynamics of early Universe, see e.g.[28]. Particularly, within the interacting vacuum energy, thenucleosynthesis sets strong constraints on the strength of thecoupling, which becomes much smaller than the unity [70].

Following previous works [59,71–74] we introduce nor-malized densities (dimensionless, positive quantities)

�A = ρA

ρcr, (12)

where ρcr = 3H2/κ2 is the critical energy density. On theone hand, the first Friedmann equation is a constraint

�r + �� + �dm + �b = 1, (13)

or

�r + �� + �m = 1, (14)

where �m ≡ �dm +�b. Because of the constraint, there areeither two or three independent normalized densities depend-ing on the interacting model. In particular, in scenario I thereare three, �r ,��,�b, in contrast, in the scenario II thereare two, namely �r ,��, with the third one being �m =1 − �r − ��, while the fourth �dm = 1 − �r − �� − �b.On the other hand, the second Friedmann equation takes theform

− H

H2 = 3

2(1 + wT ), (15)

where we have defined the total equation-of-state parameterwT = pT /ρT , which is given by

wT =∑

A

wA�A = −�� + �r

3. (16)

Finally, instead of cosmological time t we introduce the num-ber of e-folds N ≡ ln(a), and we define the time derivativesfor any quantity A as follows

A = d A

dt, (17)

A′ = d A

dN, (18)

A = H A′. (19)

Using the definitions and the cosmological equations one canobtain first order differential equations for �A with respect toN . The equations for �r ,�b are the same in all the scenariossince they are non-interacting components

�′r = �r (−1 + �r − 3��), (20)

�′b = �b(�r − 3��). (21)

123

Page 4: Dynamical systems methods and statender ... - Springer

286 Page 4 of 14 Eur. Phys. J. C (2020) 80 :286

The equation for �� depends on the interaction term Q, andtherefore there are three cases

�′� =

⎧⎪⎪⎪⎨

⎪⎪⎪⎩

3[��

(1 + �r

3 − ��

)

−νdm(1 − �r − �� − �b)], for Model I

3��

(1 + �r

3 − �� − ν�

). for Model II

(22)

Finally, q and r are computed to be

q = −1 + 3

2

(1 + �r

3− ��

), (23)

r = −q ′ + 3q

(1 + �r

3− ��

). (24)

while s can be computed using its definitions once q and rare known. Thus, we can compute the statefinder parameters,{r, s}, as a function of the red-shift z ≡ a0/a − 1 (with a0

being the present value of the scale factor a), after solvingthe system of differential equations given by Eqs. (20)–(26)in two different models for the dimensionless densities �A.Although a numerical integration of the cosmological equa-tions to obtain {r, s} is possible, in the following we willobtain exact analytical expressions, see next section.

3 Analytical solutions

The system of coupled equations may be directly integratedto obtain concrete expressions for the energy densities interms of the scale factor, as was done e.g. in Ref. [46].Although these solutions were previously reported, neitherthe statefinder diagnostic nor the phase space were analysed.This is precisely the goal of the present article, filling thusa gap in the literature. In this paper we want to completeanalysis by including the statefinder diagnostic showing, infigures, how the set {r, s} evolves for different values of red-shift, as well as the phase space of the above parameters. Westart by considering the corresponding dark matter densityρdm and dark energy density ρ� respect to the scale factorfor each model, i.e.:

ρdm ={

ρ0dma

−3(1−νdm), for Model I

ρ0dma

−3 + ν�

1−ν�ρ0

(a−3ν� − a−3

). for Model II

(25)

ρ� ={

ρ0� + νdm

1−νdmρ0

dm

(a−3(1−νdm) − 1

), for Model I

ρ0�a

−3ν� . for Model II

(26)

Thus, for each particular model the above profile densitiesgive the evolution of dark matter and dark energy respec-tively. It is important to point out that the models analysedhere boil down to the �CDM model when νi → 0. Finally,for convenience, we introduce the dimensionless Hubble rate

E(z) ≡ H(z)/H0, where H0 = 100 hkm/(Mpc s) is theHubble constant. Accordingly, the parameters {q, r, s} arecomputed as follows

q(z) = −1 + (1 + z)Ez(z)

E(z), (27)

r(z) = q(z)(1 + 2q(z)) + (1 + z)qz(z), (28)

and s(z) is given by (2), where Xz ≡ dX/dz for any quan-tity X . Using the expressions for the energy densities shownbefore, one can obtain exact analytical expressions for allquantities of interest versus red-shift, E(z), q(z), r(z), s(z),see Sect. 5.

4 Dynamical systems methods

We briefly review the stability analysis based on the natureof the fixed points (FPs), see e.g. [59,71–74]. Suppose thatfor a dynamical system with a two-dimensional phase space(x, y), its time evolution is determined by the following sys-tem of coupled first order differential equations

dx

dt= F(x(t), y(t)), (29)

dy

dt= G(x(t), y(t)). (30)

First, the fixed point(s) is (are) computed setting dx/dt =0 = dy/dt , and one has to solve the system of two algebraicequations F(x0, y0) = 0 = G(x0, y0). Then, to determinethe nature of the fixed point(s) we linearise the equationsaround that point, x(t) = x0 + δx, y(t) = y0 + δy ignoringhigher order terms. One obtains a system of two coupledlinear equations of the form

X = AX, (31)

where the column X contains the two functions δx(t), δy(t),while A is a two-dimensional matrix, the elements of whichare given by

A11 = Fx (x0, y0), (32)

A12 = Fy(x0, y0), (33)

A21 = Gx (x0, y0), (34)

A22 = Gy(x0, y0). (35)

Finally, we compute the eigenvalues λ1, λ2 of A, the sign ofwhich determines the nature of the fixed point(s). In partic-ular, the critical point is stable (A) when both eigenvaluesare negative, unstable (R) when both eigenvalues are posi-tive, and a saddle point (S) if the eigenvalues are of oppositesign. Furthermore, if q(x0, y0) < 0, wT (x0, y0) < −1/3,the fixed point at hand corresponds to acceleration. The pro-

123

Page 5: Dynamical systems methods and statender ... - Springer

Eur. Phys. J. C (2020) 80 :286 Page 5 of 14 286

Table 1 Fixed points of model IFixed point (�r , ��,�b) Eigenvalues wT q

I.a (0, 0, 1) 3, −1, 3νdm 0 1/2

I.b (1, 0, 0) 4, 1, 1 + 3νdm 1/3 1

I.c (0, 1, 0) −4, −3, −3(1 − νdm) −1 −1

I.d (0, νdm, 0) −1 − 3νdm, 3(1 − νdm), −3νdm −νdm (1 − 3νdm)/2

Table 2 Nature of fixed pointsof model I

Fixed point Existence Acceleration Nature

(0, 0, 1) ∨ νdm No S

(1, 0, 0) ∨ νdm No S (νdm < −1/3), R (νdm > −1/3)

(0, 1, 0) ∨ νdm ∨ νdm S (νdm > 1), A (νdm < 1)

(0, νdm, 0) 0 < νdm < 1 νdm > 1/3 S

Table 3 Fixed points of modelII

Fixed point (�r , ��) Eigenvalues wT q

II.a (0, 0) −1, 3(1 − ν�) 0 1/2

II.b (1, 0) 1, 4 − 3ν� 1/3 1

II.c (0, 1 − ν�) −4 + 3ν�, −3(1 − ν�) −1 + ν� −1 + 3ν�/2

cedure may be easily generalized in a straightforward mannerfor a three-dimensional phase-space.

The fixed points and their nature (stability conditions) forall two models considered in this work are shown in theTables 1, 2, 3, and 4.

4.1 Model I

In this case we obtain four critical points, which are shownin Tables 1 and 2. Point I.a is a matter dominated solutionrepresenting ordinary baryonic matter, such that �b = 1, andwith wT = 0. The eigenvalues for this critical point are

μ1 = 3, μ2 = −1, μ3 = 3νdm, (36)

and therefore it is always a saddle point. This fixed pointis not physical because it represents an era dominated bybaryons. It is well known that cold dark matter constitutesthe dominant component during the matter-dominated era atredshift 1 � z � 103, and thus during this epoch it providesthe main contribution for structure formation in the universe.

Point I.b corresponds to a radiation dominated solution,�r = 1, for which one has that wT = 1/3 and thereforethere is not acceleration. For this fixed point we find the

eigenvalues

μ1 = 4, μ2 = 1, μ3 = 1 + 3νdm, (37)

which means that it is always an unstable FP for νdm > 0.On the other hand, point I.c is a de Sitter-dominated solu-

tion for which �� = 1 and wDE = wT = −1. So, thissolution presents accelerated expansion for all values of νdm.In this case, we find the eigenvalues

μ1 = −4, μ2 = −3, μ3 = −3(1 − νdm). (38)

Clearly, for νdm < 1, point I.c is a stable node and thereforean attractor.

The last solution for this model is the fixed point I.d whichis a scaling solution with �� = νdm, as the physical require-ment implies 0 < νdm < 1. Also, this solution is charac-terized by wT = −νdm, with the decelerating and acceler-ating regimes satisfying 0 < νdm < 1/3 and νdm > 1/3,respectively. Point I.d behaves as a dark matter solution inthe limit νdm � 1, with a small contribution of dark energyproportional to νdm, during the matter dominated epoch, andthus suppressing the growth of matter perturbations. Stabilityanalysis leads us to the eigenvalues

μ1 = −1 − 3νdm, μ2 = 3 (1 − νdm) , μ3 = −3νdm. (39)

Table 4 Nature of fixed pointsof model II

Fixed point Existence Acceleration Nature

(0, 0) ∨ ν� No S (ν� < 1), A (ν� > 1)

(1, 0) ∨ ν� No R (ν� < 4/3), S (ν� > 4/3)

(0, 1 − ν�) 0 < ν� < 1 ν� < 2/3 A

123

Page 6: Dynamical systems methods and statender ... - Springer

286 Page 6 of 14 Eur. Phys. J. C (2020) 80 :286

Since we require 0 < νdm < 1, the point I.d is always asaddle point.

The critical point I.d is a dark matter dominated solutionfor the model I, with a small contribution from dark energydensity given by �� = νdm � 1 and total equation of statewT = −νdm ≈ 0. Although this fixed point can provideaccelerated expansion for νdm > 1/3, the observational con-straints on νdm � 1 [70] do not allow that this happens. So,the thermal history of the Universe is successfully reproducedfor model I provided it satisfies the restriction νdm � 1.

It is also important to note that in the present frame-work of dynamical systems the negative values for νdm areexcluded since this would imply a negative energy density�� = νdm < 0. So, as several authors usually do, we havegiven preference to maintain the physical condition ρ ≥ 0 inagreement with the weak energy condition (WEC) [28].

4.2 Model II

Model II, has three critical points which are shown inTables 3 and 4. Point II.a is a dark matter dominated solu-tion for the which �m = 1, and wT = 0. The eigenvaluesassociated with this critical point are

μ1 = −1 μ2 = 3 (1 − ν�) , (40)

and hence, one can see that it is always a saddle point for thevalue 0 < ν� < 1.

On the other hand, point II.b is a solution for which radia-tion component is the dominant one, being that �r = 1 andwT = 1/3. From stability analysis we obtain the eigenvalues

μ1 = 1 μ2 = 4 − 3ν�, (41)

which means that it is an unstable node for ν� < 4/3 and asaddle in the opposite case ν� > 4/3.

Finally, the point II.c is a scaling solution with �� =1−ν� that, interestingly, can also be an attractor with wT =−1 + ν�, allowing to alleviate the so-called cosmologicalcoincidence problem [28]. The physical requirement 0 <

�� < 1 implies 0 < ν� < 1, and from the constraintwT < −1/3, the accelerated expansion occurs for the valueν� < 2/3. For this critical point one finds the eigenvalues

μ1 = −4 + 3ν�, μ2 = −3 (1 − ν�) , (42)

So, point II.c is a stable FP for ν� < 4/3 and a saddle pointfor 4/3 < ν� < 1. Like model I, the present model II isalso physically viable to successfully reproduce the thermalhistory of the Universe from the radiation dominated era,going through the standard matter dominated era, to late timeswhen the dark energy component dominates the total energydensity and pressure of the Universe.

Considering the observational constraints as, for instance,those found in [75,76] (and references therein) the value ofthe strength of the coupling is quite smaller that unity, such

that ν� � 0.01, and this is in agreement with our theo-retical bounds. So, the results obtained from the dynamicalanalysis in this section should be supplemented by the obser-vational bounds. In fact, although the scaling solution formodel II allows us to adjust �0

� 0.73 and �0m = 0.27 for

ν� = 0.3, actually this solution cannot reached at z = 0, butonly asymptotically to reproduce the whole thermal historyof the universe. In other words, in order to obtain the physicaltrajectory in the phase space consistently with observationaldata, this scaling solution only can be reached at the future,in such a way that we need to choice smaller values of ν�

(� 10−2), allowing to have �0� 0.73 at z = 0, and asymp-

totically �� = 1 − ν� 1 with wT = 1 − ν� −1 forz → −1.

5 Statefinder analysis

In the present analysis we have calculated for both modelsI, and II, the analytical expression of the Hubble parameterE(z) as explicit function of the red-shift z, and then we haveobtained the corresponding functions q(z), r(z) and s(z). Itis important to observe that in the s−r plane, the flat �CDMscenario correspond to the point (0, 1), while that in the q−rplane the point (−1, 1) is the asymptotic de Sitter solution[77]. Since we are only focused on the evolution at late times,particularly in the transition from matter dominated era to thepresent time, we neglect the radiation component in the com-putation of E(z) and {q, r, s}. However, in order to producethe plots shown in Fig. 1 we take into account the contributioncoming from radiation.

For Model I we find

E(z) =[

�0dm

[(z + 1)3(1−νdm) − 1

]

1 − νdm

+�0b

((z + 1)3 − 1

)+ 1

] 12, (43)

and thus, for the state-finder parameters we have

q(z) = −1 +

⎢⎢⎣

3(νdm − 1)

(1 + �0

b(z+1)3νdm

�0dm

)

2

⎥⎥⎦

×

⎢⎢⎢⎢⎣

1 +(νdm−1)

[�0bz(z(z+3)+3)+1

]

�0dm

(z + 1)3(1−νdm)− 1

⎥⎥⎥⎥⎦

−1

, (44)

r(z) = 1 +[

9(1 − νdm)νdm

2

]

123

Page 7: Dynamical systems methods and statender ... - Springer

Eur. Phys. J. C (2020) 80 :286 Page 7 of 14 286

Fig. 1 The figures show the parameter space q−r and s−r for the twofirst models. To show the impact of the parameter νi = {νdm, ν�}, wecompute the functions for two different values of it, i.e.: (ii) νi = 0.01(first column), and finally (iii) νi = 0.001 (second column). In addition,

the color code is as follow: (i) solid black line correspond to �CDM,(ii) dashed red line correspond to first model and finally, (iii) dot-dashedorange line correspond to second model

×⎡

⎢⎣1 + (νdm−1)(�0

bz(z(z+3)+3)+1)

�0dm

(z + 1)3(1−νdm)− 1

⎥⎦

−1

, (45)

s(z) = (1 − νdm)νdm

νdm + (νdm−1)(�0b−1)−�0

dm�0

dm(z+1)3(1−νdm)

. (46)

Now, the Hubble rate at z � 1 yields

H ∼ 2

3t, (47)

while the parameters q, r and s behaves as

q = 1

2, (48)

r = 1, (49)

s = 1 − νdm. (50)

Therefore, in this case we recover the standard matter-dominated era with 0 < s < 1.

In the limit z → −1, the model predicts a de Sitter solutionwith

H = H0

1 − �0b + �0

dm

1 − νdm, (51)

such that q = −1, r = 1 and s = 0. Hence, at the present,z = 0, we obtain the values

q0 = −1 + 3

2

(�0

b + �0dm

)≈ −0.55, (52)

r0 = 1 − 9νdm�0dm

2, (53)

s0 = ν2�0dm

1 − �0b − �0

dm

. (54)

In calculating some numerical values we take νdm ={0.01, 0.001} for which we get q0 ≈ −0.55, r0 ≈{0.9883, 0.99883}, and s0 = {0.003714, 0.000314}.

Finally, in the case of Model II one finds

E(z) = (1 − ν�)−12

[(1 − �0

b − �0dm)(z + 1)3ν�

123

Page 8: Dynamical systems methods and statender ... - Springer

286 Page 8 of 14 Eur. Phys. J. C (2020) 80 :286

+(z + 1)3(�0b + �0

dm − ν�)]1/2

, (55)

and we also we obtain the expressions

q(z) = −1 + 3

2

⎢⎣1 − ν�(�0

b+�0dm−1)(z+1)3(ν�−1)

�0b+�0

dm−ν�

1 − (�0b+�0

dm−1)(z+1)3(ν�−1)

�0b+�0

dm−ν�

⎥⎦ , (56)

r(z) =1 + (9(ν�−1)ν�+2)(�0

b+�0dm−1)(z+1)3(ν�−1)

2(ν�−�0

b−�0dm

)

1 + (�0b+�0

dm−1)(z+1)3(ν�−1)

ν�−�0b−�0

dm

, (57)

s(z) = ν�. (58)

Similarly as in model I, for z � 1, it is recovering the stan-dard matter-dominated era, where the Hubble rate satisfiesthe relation (47). Also, in this limit we find

q = 1

2, (59)

r = 1, (60)

s = ν�, (61)

with 0 < s < 1.On the other hand, in the limit z → −1, unlike models I,

the model II behaves as a scaling solution with

H ∼ β

t, β = 2

3(1 − ν�), (62)

with a scale factor of the form a ∼ a0 (t/t0)β . It is straight-forward to check that in this limit, the statefinder parametersbecome

q = −1 + 3ν�

2, (63)

r = 1 + 9ν�

2(ν� − 1) , (64)

s = ν�. (65)

Thus, for ν� = {0.01, 0.001} one gets q0 ≈{−0.985,−0.999}, r0 ≈ {0.9555, 0.9955}, and s0 ={0.01, 0.001}.

In Fig. 1 we have depicted the behaviour of the parametersq, r and s as functions of the red-shift z, along with thetrajectories of evolution in the q − r and s − r planes, forModels I (red-dotted), and II (orange dash-dotted). For a sakeof comparison, we have also included the �CDM model(solid black line). Recall that, in order to produce the plotsshown in Fig. 1, the contribution from radiation has beentaken into account, and we also set the following values forthe fractional energy densities: �0

dm = 0.26, �0b = 0.04,

and �0r = 9 × 10−5. Naturally, �0

m ≡ �0dm + �0

b and �0� =

1 − �0m − �0

r .In these plots the behaviour of the parameters q(z), r(z)

and s(z) is in agreement with the analytical results that wehave obtained in the limit case of a negligible radiation com-ponent for z → −1. For all the three models I, and II, it is seen

Fig. 2 Two-dimensional phase-space (�m − ��) for Model II andfor ν� = 0.01. Different trajectories correspond to different initialconditions. All of them meet at the attractor ∼ (0.0, 1.0) at late timeslabelled with a point. Notice that our solution does not reach the de-Sitter solution, but it is undistinguishable from it

that the pair (s, r) starts in the left-hand side of the �CDMfixed point, which is characteristic of the hybrid expansionlaw (HEL), Chaplygin gas and Galileon models, such thats < 0 and r > 1 [78]. Let us notice that this behaviour is verydifferent from what occurs in the case of the quintessencemodel for which it is observed that the trajectory in the (s, r)plane starts in the region 0 < s < 1 and r < 1. On the otherhand, the trajectory in the (q, r) plane starts in the regionbounded by 0 < q < 1 and r > 1, being that in the caseof models I, the �CDM line is crossed at some red-shiftin the past to then evolve towards the de Sitter fixed point(q = −1, r = 1) at the future. For model II, the behaviourbecomes different because in this case the asymptotic fixedpoint is a scaling solution such that for ν� = 0.01, the bench-mark values of the fractional energy densities are �� 1and �m 0, but at z = 0, we have �� 0.7 and �m 0.3,as it has been depicted in Fig. 2.

6 Variable gravitational coupling

In this section we generalize our previous results by allow-ing a variable gravitational constant within the framework ofinteracting RVM’s.

The generalized Friedmann equations are given by

3H2 = 8πG [ρm + ρ�] , (66)

H = −4πGρm, (67)

where G(t) is a function of the cosmic time t , and ρm denotesthe non-relativistic matter energy density, including bothbaryons and dark matter. The conservation law for this model

123

Page 9: Dynamical systems methods and statender ... - Springer

Eur. Phys. J. C (2020) 80 :286 Page 9 of 14 286

can be written as

ρm + ρ� + 3Hρm + G

G(ρm + ρ�) = 0. (68)

Following our analysis for interacting RVM’s, this equationcan also be splitted into a set of two separate evolution equa-tions for the energy densities ρm and ρ�, according to Ref.[47], as follows

ρm + 3Hρm = Q, (69)

ρ� = −Q − G

G(ρ� + ρm) . (70)

Let us note that the two above equations are reduced to theEqs. (8), (9), in the case when the gravitational couplingbecomes constant. On the other hand, in order to compare thismodel with current observational data regarding dark energy,we rewrite the above Friedmann equations in the standardform

3H2 = 8πG0 [ρm + ρde] , (71)

H = −4πG0 [ρm + ρde + pde] , (72)

with G0 being is the constant gravitational coupling. In theseequations we have introduced the effective energy densityand pressure density of dark energy, defined as

ρde = −ρm + G

G0(ρm + ρ�) , (73)

pde = − G

G0ρ�, (74)

respectively. This effective dark energy density includes theeffect of both vacuum energy density ρ�(t) and the dynami-cally changing gravitational coupling G(t). It is straightfor-ward to show that it satisfies the evolution equation

ρde + 3H(ρde + pde) = −Q. (75)

Therefore, it is easy to see that the two evolution equations(70) and (75) together are consistent with the energy conser-vation law for ρm and ρDE .

As it is usually done, we introduce the fractional and crit-ical energy densities

�de = ρDE

ρcr, ρcr = 3H2

8πG0, (76)

the EOS parameter of dark energy

wde = pdeρde

= − G

G0+

(−1 + G

G0

)1

�DE, (77)

and the total EOS parameter

wT = pTρT

= −1 + (1 − �DE )G

G0. (78)

With this, the condition for having accelerated expansionbecomes we f f < −1/3, or equivalently q < 0.

6.1 An example for variable G

In order to obtain concrete results, we assume the follow-ing phenomenological ansatz for the gravitational couplingdepending on the Hubble rate, according to [47]

G(X) = G0

1 + νG ln X, (79)

where X ≡ E2 = (H/H0)2. Also, in order to extend our pre-

vious analysis for interacting RVM’s, we consider the cou-pling between DE and DM to be Q = 3νmHρm (Model II).For this coupling function Q and ansatz (79), the equations(70) and (75) take the form

dρmdX

= (1 − νm)ρ0cr (νG ln X + 1), (80)

dρdedX

= ρ0cr [νm − νG(1 − νm) ln X ] , (81)

whose solutions are given by

ρm(X) = ρ0m + ρ0

cr [−1 + νm + νG (1 − νm)]

+Xρ0cr [1 − νm − νG (1 − νm) (1 − ln X)] , (82)

and

ρde(X) = ρ0de − ρ0

cr [νm + νG (1 − νm)]

+Xρ0cr [νm + νG (1 − νm) (1 − ln X)] , (83)

respectively.Upon replacement of the above solutions into Eq. (71),

one obtains ρ0cr = ρ0

m + ρ0de. On the other hand, doing the

same but now in Eq. (72) we find

dX

dN= 3

[ρ0de/ρ

0cr − νm − (1 − νm) (νG + X (1 − νG (1 − ln X)))

]

1 + νG ln X.

(84)

After solving this equation, we obtain the dimensionlessHubble rate squared E2(N ), and then H(N ), or equivalently,H(z), by introducing N = − ln(1 + z).

Also, from (76), one has that

�de(X) = νm +[νG(νm − 1) − νm + �0

de

] 1

X+νG(1 − νm) (1 − ln X) , (85)

and �m(X) = 1 − �de(X). From Eqs. (77), (78), we find

wde(X) = − 1

1 + νG ln X[

1 + νG X ln X

νG((1 − νm)X + νm − 1)+XνG(νm−1) ln X+(X − 1)νm+�0de

],

(86)

and

123

Page 10: Dynamical systems methods and statender ... - Springer

286 Page 10 of 14 Eur. Phys. J. C (2020) 80 :286

0 1 2 3 4

50

100

150

200

250

300

350

400

z0 2 4 6 8 10

0.0

0.5

1.0

1.5

2.0

2.5

3.0

z

rH(z)

Fig. 3 It is shown the evolution of the Hubble rate H(z) (left panel)as a function of red-shift z, for the G-varying model defined in Eq.(79), with the coupling function Q = 3νdm Hρdm , representing theinteraction between dark energy and dark matter, for νm = 0.01 andνG = 1×10−3 (dashed line), as also, the Hubble rate of �CDM model,along with the Hubble data from Refs. [79], and [80]. In the right panel

it is depicted the behaviour of the exact relative difference (percentage)defined as r H(z) ≡ 100 × |H − H�CDM |/H�CDM with respect tothe concordance model, for νdm = 0.01 and two different values of νG ,the first value νG = 5 × 10−4 (short-dashed line) and the second oneνG = 1×10−3 (large-dashed line). We have used H0 = 67.4 Km/(Mpcs) from Planck 2018 [34]

wT (X) = 1

(1 + νG ln X)

[ (1 + 1

X

)

[νm + νG (1 − νm)] − �0de

X− νGνm ln X

]. (87)

The Eq. (84) is an autonomous equation which can betreated following the same framework of dynamical systems.This equation has a single critical point Xc which can bedetermined from

ln Xc = Xc(νG − 1)(νm − 1) − νGνm + νG + νm − �0de

XcνG(νm − 1).

(88)

For example, numerically we have found that for �0de

0.73, νdm 0.01 and νG 1 × 10−3, the value Xc isapproximately Xc 0.73. This result gives us the asymptotic(future) value of the Hubble rate by using the relation Hc =H0

√Xc. So, from the observational data of Planck 2018 Ref.

[34], one has that H0 = 67.4 km/(Mpc s) and therefore H0 57.48 km/(Mpc s) at z → −1.

By substituting in Eqs. (85), (85), and (85), one can seethat it is a Sitter solution representing dark energy dominancewith �de = 1, wde = −1, and wT = −1.

The stability of this fixed point can be studied by consid-ering the solution

X (N ) = Xc + δX (N ), (89)

where the perturbation δX (N ) satisfies δX (N ) � 1. Thus,replacing it in Eq. (84) we obtain

dδX

dN= −3(1 − νm)δX, (90)

whose solution is

δX = CeμN , (91)

with μ = −3(1 − νm). So, since νm � 1 then μ < 0 andaccordingly, the fixed point is always an attractor. We notethat the stability does not depend on νG at leading order inperturbation.

Similarly, by using Eqs. (28) and (2), the statefinderparameters r and s are computed to be

r = 1 − 9νm(1 − �0de)(z + 1)3(1−νm )

2X (1 + νG ln X)

−9νG(1 − �0de)

2(z + 1)6(1−νm )

2X2(1 + νG ln X)3 , (92)

s = (1 − �0de)

[νm X (1 + νG ln X)2 + νG(1 − �0

de)(z + 1)3(1−νm )]

X (1 + νG ln X)2[−1 + �0

de + X (z + 1)3(νm−1)(1 + νG ln X)] ,

(93)

which reduce to those already obtained for Model II in thelimit νG → 0.

In Fig. 3 (left panel), it is shown the evolution of the Hub-ble rate H(z) for the present model by solving the differentialequation (84) for some values of νdm and νG . It is also addedthe Hubble rate H�CDM of the �CDM model, along withthe current available data for H(z) from [79] and [80]. Also,we depict the behaviour of the exact relative difference withrespect to the concordance model, for fixed νdm and a pair ofdifferent values of νG . We take νdm = 0.01 and two differentvalues of νG , the first value νG = 5 × 10−4 and the secondone νG = 1 × 10−3, which are included into the physicalrange νG ∈ [

5 × 10−4, 1 × 10−3], obtained from observa-

tions, see for example Refs. [40,42,81]. It is observed anincreasing of r H for higher red-shifts, z � 2, and after thepresent time z = 0, in the future. Particularly, for z 10we obtain r H 3%, whereas that for z = −1 we have r H 0.18%.

In Fig. 4 we plot the variation with respect to z of some cos-mological parameters such the fractional energy densities of

123

Page 11: Dynamical systems methods and statender ... - Springer

Eur. Phys. J. C (2020) 80 :286 Page 11 of 14 286

0 2 4 6 8 100.0

0.2

0.4

0.6

0.8

1.0

z0 2 4 6 8 10

1.5

1.0

0.5

0.0

z

w

Fig. 4 In the left panel we depict the behaviour of the fractionalenergy densities of dark energy �de (solid line) and cold dark mat-ter �dm (dashed line), as functions of redshift z, for νm = 0.01 andνG = 5 × 10−4. In the right panel we show the behaviour of theEOS parameter of dark energy wde (short-dashed line) and the totalEOS parameter wT (solid line) as functions of z, for νm = 0.01 andνG = 5 × 10−4. Also, the large-dashed line corresponds to values of

wde, but for a larger value of νG , in such a way that we now haveνG = 1 × 10−3. At the present time, at z = 0, they take the values�de 0.73, �dm = 0.27, wde −1, and wT −1. Asymptotically,the model approaches to an attractor fixed point which is a de Sitterdark energy dominated solution with �de = 1, �dm = 0, wde = −1,and wT = −1

0 2 4 60.95

0.96

0.97

0.98

0.99

1.00

z

r

0 2 4 60.0

0.1

0.2

0.3

0.4

0.5

0.6

z

s

Fig. 5 It is depicted the behaviour of the statefinder parameters r (leftpanel) and s (right panel) as functions of redshift z, for νm = 0.01,and νG = 5 × 10−4 (short-dashed line) and the second νG = 1 × 10−3

(large-dashed line). Solid line represents the values of r and s for �CDM

model. It is seen that asymptotically, when z → −1, the system tendsto a de Sitter solution, r = 1 and s = 0, consistently with previousresults of dynamical systems

dark energy �de(z), dark matter �dm(z), the EOS parameterof dark energy wde(z), and the total EOS parameter wT (z).It is seen that the model can explain the current acceleratedexpansion of the universe and at z = 0 it gives �0

de 0.73,�0

dm 0.27, w0de −1, and w0

T −0.73. Also, whenz → −1, the model tends asymptotically to an attractorwhich is a de Sitter solution with �de = 1, �dm = 0,wde = −1 and wT = −1. For higher red-shifts we observethat wde becomes more sensitive to the values of νG , tak-ing smaller values than −1 and going deeper in the phantomregime for larger values of νG . Nevertheless, let us note thatthe energy density of dark energy decays very quickly andthe effective cosmic fluid behaves as nonrelativistic matterwith wT 0, and therefore allowing the existence of thestandard matter-dominated era [28,82].

Finally, in Fig. 5 we show the evolution of the statefinderparameters r(z) (left panel) and s(z) (right panel) as functionsof red-shift, for the same set of values of parameters used inabove plots. It can be seen that at z = 0, and for νdm = 0.01,νG = 5×10−4 (short dashed line), these parameters take thevalues r 0.988 and s 3.75 × 10−3. For the larger valueνG = 1×10−3 (large dashed line), at z = 0, we get r 0.988and s 3.80 ×10−3. So, for lower red-shift, there is a smalldifference (of the order of 4% for z � 2) between the resultswhen varying νG , and this difference is much smaller for thevalues of r than for s. When z → −1, the trajectory of thesystem in the plane of statefinder parameters tends towardthe de Sitter expansion at the future, with r = 1 and s = 0.

123

Page 12: Dynamical systems methods and statender ... - Springer

286 Page 12 of 14 Eur. Phys. J. C (2020) 80 :286

7 Concluding remarks

In summary, in the present work we have applied phase-spacedynamical techniques to three running vacuum dark energymodels, and we have computed the statefinder parameters asfunctions of the red-shift.

From our dynamical analysis we have shown that the twomodels, I, and II can explain the current accelerated expan-sion phase of the Universe, being that the correspondingfixed point, either a de Sitter solution (Model I) or scalingsolution (Model II), is also an attractor in all the cases, pro-vided that νi = {νdm, ν�} < 1. Also, for Model I and IIthe thermal history of the universe can be successfully repro-duced, from the radiation-dominated era, passing through thematter-dominated era, and finally reaching the dark energy-dominated phase.

More interestingly, in the case of Model I, the fixed pointrepresenting the matter era is a scaling solution in whichthe dark energy density has a contribution to the total energydensity during the dark matter era, with �� = νdm. So, whenthe parameter νdm is small this contribution is also small.Thus, the parameter νdm indicates a slight deviation withrespect to the standard mater era whose value may be moreconstrained from large-scale structure (LSS) data. It is due tothe fact that when the universe enters in this fixed point thegrowing of matter density perturbations can be suppressed bythe presence of dark energy, such that the dark matter densitycontrast grows less rapidly that the first power of the scalefactor, depending on the amount of dark energy [83]. On theother hand, in the case of model II, the final attractor is ascaling solution with accelerated expansion which satisfies�� = 1 − ν�, wDE = −1 and wT = −1 + ν�. Thisclass of solution is very interesting because it provides anatural mechanism in alleviating the fine-tuning problem, orcosmological coincidence problem of dark energy [28]. Itcan adjust the current values of the cosmological parameterssuch as �0

� = 0.7 and �0m = 0.3, and at the same time

explain the accelerated expansion.Regarding the statefinder analysis, first let us recall that

the pair {r, s} is defined using third order time derivativesof the scale factor, and that the statefinder parameters havethe potential to discriminate between different dark energymodels. Now, we should discriminate between the two casesshown in Fig. 1. The first column was plotted for νi = 0.01,the second column was depicted for νi = 0.001. Start-ing from the left panel, we observe evident discrepanciesbetween the models. This difference is also natural becauserunning vacuum models are quantum-inspired, which meansthat the deviations with respect to the classical counterpartshould be, in general, small. Taking the above idea seriously,the parameter νi , which encodes the quantum features, shouldbe taken in such a way that the effect on the classical solu-tion will be soft. We then claim that the last column in Fig. 1

should be taken as a more suitable situation. Furthermore,notice that when νi is taken to be close to zero (second col-umn), the deceleration parameter q(z) looks qualitative iden-tical to those for �CDM. If we now analyse the parameter r ,we observe once more a notorious similarity to the standardscenario, although the second parameter (i.e., s) exhibits aremarkable difference. Thus, although classically these mod-els should be equivalent, at the level of the statefinder diag-nostic, this is not the case.

Finally, it has been performed a further analysis regard-ing to a class of models with variable gravitational couplingwithin the interacting vacuum scenario, by using the toolsdynamical system and statefinder analysis. In doing so, wehave studied a particular model for G(H) from Ref. [47],and Q = 3νdmHρdm . It has been introduced an effectivecosmic fluid for describing DE by defining, both effectiveenergy density and pressure, which contain the contributionof the running of gravitational coupling. We have shown thatthe model has only fixed point which is an attractor and deSitter solution, allowing to adjust the current data of H(z),along with the other cosmological parameters such that thefractional energy density of DE and the equation of stateof dark energy at the present time. In particular, we foundthat the effect of νdm becomes more significant for higherredshift, and at the future when z → −1 in comparisonwith the �CDM model. Regarding the statefinder analysiswe observe that the s parameter becomes more sensitiveto higher redshifts than the r parameter, when we vary thecouplingνG . As a final remark, an exhaustive study includ-ing several anszats for both the gravitational coupling andthe interaction between DE and DM deserves a separatedproject, reason why these ideas will de addressed in a futurework.Note added: One day before we received the referee report, anew work related to ours appeared [84], which indicates thatthe topic is interesting. Our work was carried out in completeindependence from them, and vice versa. In that work, theauthors have analysed the dynamics and evolution of several�-varying cosmological models. The critical points and theirnature have been determined, and the corresponding phasespace is shown, although they have not discussed at all thestatefinder parameters. We have checked that where there isoverlap their results are in agreement with ours, which is aconfirmation that both calculations are error-free.

Acknowledgements We are grateful to the anonymous reviewer fora careful reading of the manuscript, for her/his constructive criticismand for valuable comments and suggestions. The author G. P. thanks theFundação para a Ciência e Tecnologia (FCT), Portugal, for the financialsupport to the Center for Astrophysics and Gravitation-CENTRA, Insti-tuto Superior Técnico, Universidade de Lisboa, through the Grant no.UID/FIS/00099/2013. The author Á. R. acknowledges DI-VRIEA forfinancial support through Proyecto Postdoctorado 2019 VRIEA-PUCV.The author N. V. was supported by Comisión Nacional de Ciencias yTecnología of Chile through FONDECYT Grant no. 11170162. Addi-

123

Page 13: Dynamical systems methods and statender ... - Springer

Eur. Phys. J. C (2020) 80 :286 Page 13 of 14 286

tionally, N. V. would like to express his gratitude to the Instituto SuperiorT écnico of Universidade de Lisboa for its kind hospitality during thefinal stages of this work. The author G. O acknowldeges DI-VRIEA forfinancial support through Proyecto Postdoctorado 2019 VRIEA-PUCV.

Data Availability Statement This manuscript has no associated data orthe data will not be deposited. [Authors’ comment: The present work isa theoretical study, and therefore no experimental data has been listed.]

Open Access This article is licensed under a Creative Commons Attri-bution 4.0 International License, which permits use, sharing, adaptation,distribution and reproduction in any medium or format, as long as yougive appropriate credit to the original author(s) and the source, pro-vide a link to the Creative Commons licence, and indicate if changeswere made. The images or other third party material in this articleare included in the article’s Creative Commons licence, unless indi-cated otherwise in a credit line to the material. If material is notincluded in the article’s Creative Commons licence and your intendeduse is not permitted by statutory regulation or exceeds the permit-ted use, you will need to obtain permission directly from the copy-right holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.Funded by SCOAP3.

References

1. A.G. Riess et al., Astron. J. 116, 1009 (1998)2. S. Perlmutter et al., Astrophys. J. 517, 565 (1999)3. W.L. Freedman, M.S. Turner, Rev. Mod. Phys. 75, 1433 (2003).

arXiv:astro-ph/03084184. A. Einstein, Ann. Phys. 49, 769–822 (1916)5. A. Einstein, Sitzungsber. Preuss. Akad. Wiss. Berlin (Math. Phys.)

1917, 142 (1917)6. S. Weinberg, Rev. Mod. Phys. 61, 1 (1989)7. Y.B. Zeldovich, JETP Lett. 6, 316 (1967). (Pisma Zh. Eksp. Teor.

Fiz. 6, 883, 1967)8. J. Garriga, A. Vilenkin, Phys. Rev. D 64, 023517 (2001)9. T. Padmanabhan, H. Padmanabhan, Int. J. Mod. Phys. D 22,

1342001 (2013)10. A. Mikovic, M. Vojinovic, EPL 110(4), 40008 (2015)11. F. Canales, B. Koch, C. Laporte, Á. Rincón, arXiv:1812.10526

[gr-qc]12. T.P. Sotiriou, V. Faraoni, Rev. Mod. Phys. 82, 451 (2010).

arXiv:0805.1726 [gr-qc]13. A. De Felice, S. Tsujikawa, Living Rev. Relat. 13, 3 (2010).

arXiv:1002.4928 [gr-qc]14. W. Hu, I. Sawicki, Phys. Rev. D 76, 064004 (2007).

arXiv:0705.1158 [astro-ph]15. A.A. Starobinsky, JETP Lett. 86, 157 (2007)16. D. Langlois, Prog. Theor. Phys. Suppl. 148, 181 (2003).

arXiv:hep-th/020926117. R. Maartens, Living Rev. Relat. 7, 7 (2004). arXiv:gr-qc/031205918. G.R. Dvali, G. Gabadadze, M. Porrati, Phys. Lett. B 485, 208

(2000). arXiv:hep-th/000501619. C. Brans, R.H. Dicke, Phys. Rev. 124, 925 (1961)20. C.H. Brans, Phys. Rev. 125, 2194 (1962)21. J.C.B. Sanchez, L. Perivolaropoulos, Phys. Rev. D 81, 103505

(2010). arXiv:1002.2042 [astro-ph.CO]22. G. Panotopoulos, Á. Rincón, Eur. Phys. J. C 78(1), 40 (2018).

arXiv:1710.02485 [astro-ph.CO]23. B. Ratra, P.J.E. Peebles, Phys. Rev. D 37, 3406 (1988)24. I.Y. Aref’eva, A.S. Koshelev, S.Y. Vernov, Theor. Math.

Phys. 148, 895 (2006). (Teor. Mat. Fiz. 148, 23, 2006).arXiv:astro-ph/0412619

25. R. Lazkoz, G. Leon, Phys. Lett. B 638, 303 (2006).arXiv:astro-ph/0602590

26. J.S. Bagla, H.K. Jassal, T. Padmanabhan, Phys. Rev. D 67, 063504(2003). arXiv:astro-ph/0212198

27. C. Armendariz-Picon, V.F. Mukhanov, P.J. Steinhardt, Phys. Rev.D 63, 103510 (2001). arXiv:astro-ph/0006373

28. E.J. Copeland, M. Sami, S. Tsujikawa, Int. J. Mod. Phys. D 15,1753 (2006). arXiv:hep-th/0603057

29. B. Ryden, Nat. Phys. 13(3), 314 (2017)30. L. Verde, P. Protopapas, R. Jimenez, Phys. Dark Univ. 2, 166

(2013). arXiv:1306.6766 [astro-ph.CO]31. K. Bolejko, Phys. Rev. D 97(10), 103529 (2018)32. E. Mrtsell, S. Dhawan, arXiv:1801.07260 [astro-ph.CO]33. P.A.R. Ade et al. [Planck Collaboration], Astron. Astrophys. 594,

A13 (2016). arXiv:1502.01589 [astro-ph.CO]34. N. Aghanim et al. [Planck Collaboration], arXiv:1807.06209

[astro-ph.CO]35. A.G. Riess et al., Astrophys. J. 826(1), 56 (2016).

arXiv:1604.01424 [astro-ph.CO]36. A.G. Riess et al., Astrophys. J. 861(2), 126 (2018).

arXiv:1804.10655 [astro-ph.CO]37. E. Mrtsell, S. Dhawan, JCAP 1809(09), 025 (2018).

arXiv:1801.07260 [astro-ph.CO]38. E. Macaulay, I.K. Wehus, H.K. Eriksen, Phys. Rev. Lett. 111(16),

161301 (2013). arXiv:1303.6583 [astro-ph.CO]39. S. Basilakos, S. Nesseris, Phys. Rev. D 96(6), 063517 (2017).

arXiv:1705.08797 [astro-ph.CO]40. J. Sola, A. Gomez-Valent, J. de Cruz Pérez, Astrophys. J. 811, L14

(2015). arXiv:1506.05793 [gr-qc]41. S. Basilakos, J. Sola, Phys. Rev. D 92(12), 123501 (2015).

arXiv:1509.06732 [astro-ph.CO]42. J. Sola, A. Gómez-Valent, J. de Cruz Pérez, Astrophys. J. 836(1),

43 (2017). arXiv:1602.02103 [astro-ph.CO]43. J. Sola, A. Gómez-Valent, J. de Cruz Pérez, Int. J. Mod. Phys. A

32(19–20), 1730014 (2017). arXiv:1709.07451 [astro-ph.CO]44. J. Sola, A. Gómez-Valent, J. de Cruz Pérez, Phys. Lett. B 774, 317

(2017). arXiv:1705.06723 [astro-ph.CO]45. A. Gomez-Valent, J. Sola, EPL 120(3), 39001 (2017).

arXiv:1711.00692 [astro-ph.CO]46. J Sola Peracaula, Int. J. Mod. Phys. A 33(31), 1844009 (2018)47. H. Fritzsch, J. Sola, R.C. Nunes, Eur. Phys. J. C 77, 193 (2017)48. M. Reuter, H. Weyer, Phys. Rev. D 69, 104022 (2004)49. B. Koch, I.A. Reyes, Á. Rincón, Class. Quantum Gravity 33(22),

225010 (2016)50. A. Hernández-Arboleda, Á. Rincón, B. Koch, E. Contreras, P. Bar-

gueño, arXiv:1802.05288 [gr-qc]51. A.G. Riess, S. Casertano, W. Yuan, L.M. Macri, D. Scolnic, Astro-

phys. J. 876(1), 85 (2019)52. G. Risaliti, E. Lusso, Nat. Astron. 3(3), 272 (2019)53. S. Kumar, R.C. Nunes, Phys. Rev. D 94(12), 123511 (2016)54. S. Kumar, R.C. Nunes, Phys. Rev. D 96(10), 103511 (2017)55. W. Yang, S. Pan, E. Di Valentino, R.C. Nunes, S. Vagnozzi, D.F.

Mota, JCAP 1809, 019 (2018)56. S. Kumar, R.C. Nunes, Eur. Phys. J. C 77(11), 734 (2017)57. W. Yang, S. Vagnozzi, E. Di Valentino, R.C. Nunes, S. Pan, D.F.

Mota, JCAP 1907, 037 (2019)58. B. Wang, E. Abdalla, F. Atrio-Barandela, D. Pavon, Rep. Prog.

Phys. 79(9), 096901 (2016)59. G. Kofinas, G. Panotopoulos, T.N. Tomaras, JHEP 0601, 107

(2006). arXiv:hep-th/051020760. V. Sahni, T.D. Saini, A.A. Starobinsky, U. Alam, JETP Lett.

77, 201 (2003). (Pisma Zh. Eksp. Teor. Fiz. 77, 249, 2003).arXiv:astro-ph/0201498

61. U. Alam, V. Sahni, T.D. Saini, A.A. Starobinsky, Mon. Not. R.Astron. Soc. 344, 1057 (2003). arXiv:astro-ph/0303009

62. J. Albert et al. [SNAP Collaboration], arXiv:astro-ph/0507458

123

Page 14: Dynamical systems methods and statender ... - Springer

286 Page 14 of 14 Eur. Phys. J. C (2020) 80 :286

63. J. Albert et al. [SNAP Collaboration], arXiv:astro-ph/050745964. W. Zimdahl, D. Pavon, Gen. Relat. Gravit. 36, 1483 (2004).

arXiv:gr-qc/031106765. X. Zhang, Phys. Lett. B 611, 1 (2005). arXiv:astro-ph/050307566. Px Wu, Hw Yu, Int. J. Mod. Phys. D 14, 1873 (2005).

arXiv:gr-qc/050903667. B. Chang, H. Liu, L. Xu, C. Zhang, Mod. Phys. Lett. A 23, 269

(2008). arXiv:0704.3670 [astro-ph]68. G. Panotopoulos, Nucl. Phys. B 796, 66 (2008). arXiv:0712.1177

[astro-ph]69. E.L.D. Perico, D.A. Tamayo, JCAP 1708, 026 (2017)70. C. Espana-Bonet, P. Ruiz-Lapuente, I.L. Shapiro, J. Sola, JCAP

0402, 006 (2004). arXiv:hep-ph/031117171. M. Goliath, G.F.R. Ellis, Phys. Rev. D 60, 023502 (1999).

arXiv:gr-qc/981106872. E.J. Copeland, A.R. Liddle, D. Wands, Phys. Rev. D 57, 4686

(1998). arXiv:gr-qc/971106873. C.G. Boehmer, G. Caldera-Cabral, R. Lazkoz, R. Maartens, Phys.

Rev. D 78, 023505 (2008). arXiv:0801.1565 [gr-qc]74. L Lopez Honorez, O. Mena, G. Panotopoulos, Phys. Rev. D 82,

123525 (2010). arXiv:1009.5263 [astro-ph.CO]

75. S. Kumar, R.C. Nunes, S.K. Yadav, Eur. Phys. J. C 79, 576 (2019).arXiv:1903.04865 [astro-ph.CO]

76. E. Di Valentino, A. Melchiorri, O. Mena, S. Vagnozzi,arXiv:1908.04281 [astro-ph.CO]

77. L.N. Granda, Mod. Phys. Lett. A 28, 1350117 (2013).arXiv:1308.6565 [gr-qc]

78. Ö. Akarsu, S. Kumar, R. Myrzakulov, M. Sami, L. Xu, JCAP 1401,022 (2014). arXiv:1307.4911 [gr-qc]

79. X.L. Meng, X. Wang, S.Y. Li, T.J. Zhang, arXiv:1507.02517 [astro-ph.CO]

80. O. Farooq, B. Ratra, Astrophys. J. 766, L7 (2013). arXiv:1301.5243[astro-ph.CO]

81. J Sola Peracaula, J. de Cruz Pérez, A. Gómez-Valent, EPL 121,39001 (2018). arXiv:1606.00450 [gr-qc]

82. M. Gonzalez-Espinoza, G. Otalora, J. Saavedra, N. Videla, Eur.Phys. J. C 78, 799 (2018). arXiv:1808.01941 [gr-qc]

83. L. Amendola, Phys. Rev. D 62, 043511 (2000).arXiv:astro-ph/9908023

84. G. Papagiannopoulos, P. Tsiapi, S. Basilakos, A. Paliathanasis,arXiv:1911.12431 [gr-qc]

123