Next Article in Journal
The Low-Energy Module (LEM): Development of a CubeSat Spectrometer for Sub-MeV Particles and Gamma-Ray Burst Detection
Previous Article in Journal
Stacking Lake Ice Analog Cryotectonic Dynamics on Dione’s Wispy Terrain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Forecasts for ΛCDM and Dark Energy Models through Einstein Telescope Standard Sirens †

by
Matteo Califano
1,2,*,
Ivan de Martino
3,
Daniele Vernieri
1,2,4 and
Salvatore Capozziello
1,2,4
1
Scuola Superiore Meridionale, Largo San Marcellino 10, I-80138 Napoli, Italy
2
INFN Sezione di Napoli, Complesso Universitario di Monte Sant’ Angelo, Edificio G, Via Cinthia, I-80126 Napoli, Italy
3
Departamento de Fisica Fundamental, Universidad de Salamanca, P. de la Merced S/N, 37008 Salamanca, Spain
4
Dipartimento di Fisica, Università di Napoli “Federico II”, Complesso Universitario di Monte Sant’ Angelo, Edificio G, Via Cinthia, I-80126 Napoli, Italy
*
Author to whom correspondence should be addressed.
Presented at the 2nd Electronic Conference on Universe, 16 February–2 March 2023; Available online: https://ecu2023.sciforum.net/.
Phys. Sci. Forum 2023, 7(1), 20; https://doi.org/10.3390/ECU2023-14032
Published: 16 February 2023
(This article belongs to the Proceedings of The 2nd Electronic Conference on Universe)

Abstract

:
Gravitational wave (GW) astronomy provides an independent way to estimate cosmological parameters. The detection of GWs from a coalescing binary allows a direct measurement of its luminosity distance, so these sources are referred to as “standard sirens” in analogy to standard candles. We investigate the impact of constraining cosmological models on the Einstein Telescope, a third-generation detector which will detect tens of thousands of binary neutron stars. We focus on non-flat ΛCDM cosmology and some dark energy models that may resolve the so-called Hubble tension. To evaluate the accuracy down to which ET will constrain cosmological parameters, we consider two types of mock datasets depending on whether or not a short gamma-ray burst is detected and associated with the gravitational wave event using the THESEUS satellite. Depending on the mock dataset, different statistical estimators are applied: one assumes that the redshift is known, and another marginalizes it, taking a specific prior distribution.

1. Introduction

The observation of GWs from the coalescence of merging binary black holes (BBH) [1,2] and binary neutron stars (BNS) [3] gives an alternative tool to test general relativity, relativistic astrophysics, and cosmology. We usually refer to GWs as “standard sirens” because, in analogy to standard candles, they bring direct information on the luminosity distance of sources [4,5].
Contrary to most common electromagnetic (EM) distance measurements, the distance estimate with GWs is an absolute measurement. Hence, standard sirens do not rely on the so-called cosmic distance ladder. Therefore, they are free from possible systematics arising from the calibration on other cosmic distance indicators.
In the Friedmann–Robertson–Walker cosmology, the most general form of the distance–redshift relation reads [6]:
d L ( z ) =   c H 0 1 + z Ω k , 0   sin h [   Ω k , 0       0 z d z E ( z ) ] ,  
where c is the speed of light, H 0 is the Hubble constant, Ω k , 0   is the normalized energy density of the spatial curvature of the Universe, and E(z) is a function of redshift, which in general depends on all the cosmological parameters that describe the background expansion of the Universe in any given cosmological model. The data ( d L , z )   allow us to constrain the cosmological parameters in the distance–redshift relation. In particular, one can infer the Hubble constant H 0 to the leading order, and beyond that the dark matter and dark energy fractions Ω m , Ω Λ   of Λ CDM cosmology, or the dark-energy (DE) equation-of-state parameters.
Although GWs offer an alternative method to obtain distances in cosmology, they are not free of issues. In particular, the redshift parameter in the waveform is completely degenerate with the system masses. We can break the degeneracy by extrapolating the information on the redshift from an electromagnetic signal. The main techniques are based on the statistical identification of the host galaxy of the GW source [4,7] or the seeking of electromagnetic emissions following the GWs, such as short gamma-ray burst (GRB) [3]. Another possibility relies on assuming the redshift probability distribution of GW events known from population synthesis simulations [8,9].
Nowadays, the LIGO/Virgo/KAGRA collaboration best estimation of the Hubble constant is H 0 = 68 8 + 6 km s 1   Mpc 1 , at a 68% of confidence level with the statistical identification of the host galaxy [7]. However, so far, the GWs do not help solve the so-called Hubble tension because the accuracy is still too high, and the estimations agree with both the late-time and the early-time measurements [10,11,12,13].
Nevertheless, the next generation of GW detectors, e.g., the Einstein Telescope (ET), will offer the possibility to achieve an accuracy of the Hubble constant below 1% [14]. Here, we will focus on the simulation ET standard sirens. Moreover, we assume that the redshift of the coincident short GRB will be detected using the Transient High Energy Sources and Early Universe Surveyor (THESEUS) [15,16,17]. We forecast the accuracy of cosmological parameters for a non-flat ΛCDM and a set of DE models introduced to solve the Hubble tension [18,19]. We consider the following parametrizations of the E(z) function:
  • Non-flat Λ CDM, with the E(z) function defined by [6]
E 2 ( z ) = Ω m , 0 ( 1 + z ) 3 + Ω k , 0 ( 1 + z ) 2 + Ω Λ , 0   ;
  • Non-flat ω CDM, with the E(z) function defined by [20]
E 2 ( z ) = Ω m , 0 ( 1 + z ) 3 + Ω k , 0 ( 1 + z ) 2 + Ω Λ , 0 ( 1 + z ) 3 ( 1 + ω DE )   ;
E 2 ( z ) = Ω m , 0 ( 1 + z ) 3 + Ω Λ , 0 [ ( 1 + z ) 3 ( 1 + ω DE eff ) + ξ 3 ω DE eff ( 1 ( 1 + z ) 3 ω DE eff ) ( 1 + z ) 3 ] ,
where ω DE eff = ω D E + ξ / 3 and ξ is the coupling constant;
  • Time-Varying Gravitational Constant, [26]
E 2 ( z ) = Ω m , 0 ( 1 + z ) ( 3 δ G ) + Ω Λ , 0 ( 1 + z ) δ G   Ω m , 0 Ω Λ , 0   ,  
with δ G representing the parametrization of Gravitational Constant evolution;
E 2 ( z ) = Ω m , 0 ( 1 + z ) 3 + Ω Λ , 0 [ 1 tanh ( Δ log 10 ( 1 + z 1 + z t ) ) 1 + tanh ( Δ log 10 ( 1 + z t ) ) ] ,
where Δ is a free parameter and z t is the epoch where the matter energy density and the DE density are equal.
In the following sections, we briefly summarize the procedure used to build up the mock data catalog (Section 2) and the statistical analysis techniques (Section 3). Finally, in Section 2, we discuss our results.

2. Mock Data Generation

Following the procedure illustrated in [30,31], we simulated the GW events to forecast the precision down to which ET would be able to constrain the cosmological parameters. We wanted to consider only the BNS mergers because we could detect their EM counterpart. To generate the synthetic dataset, we assumed, as a fiducial cosmological model, a Λ CDM with best-fit values given by [13], which were H 0 = 67.66 km s 1   Mpc 1 , Ω m , 0 = 0.31 , Ω Λ , 0 = 0.69 and Ω k , 0 = 0.0 . Then, we extracted the redshift of the source from a probability distribution, p(z), defined from the star formation rate (SFR) and the time delay distribution. The function p(z) is
p ( z ) = N R m ( z ) 1 + z dV ( z ) dz
where N is a normalization factor, dV(z)/dz is the comoving volume element, and R m ( z ) is the merger rate per unit of volume in the source frame. We can parametrize the rate R m ( z )   in terms of the SFR R f ( z ) [32], and the time delay distribution P ( t d ) t d 1 as suggested by population synthesis models [33].
Then, using the latest power spectral density of ET, we simulated the detector response to estimate the number and the parameters of GW events. Hence, we selected the events above a given of the signal-to-noise ratio (SNR). We adopted a SNR threshold equal to 9. Finally, we added a Gaussian noise component, N ( d L fid , σ d L ) to our estimations of the luminosity distances d L fid . based on the fiducial cosmological model. The variance counts for different sources of uncertainties:
σ d L = σ inst 2 + σ lens 2 + σ pec 2
The first term is the most relevant due to the instrumental part. At leading order, σ inst is strictly related to the SNR through the relation σ inst = 2 d L / SNR [34]. The second and the last ones are related to some extra contributions in the noise due to the observational features. We considered the lensing [35] and the peculiar velocity of the host galaxy contribution [36]. Setting a duty cycle for ET equal to 80%, we built our mock catalogs containing GWs events for one, five, and ten years of observational runs. We estimated a rate of 0.5 × 10 4 events per year.
Since the number of combined events is strictly affected by the features of the satellite, we had to set the duty cycle of the THESEUS satellite to 80% [15] and the sky coverage to 1/2. Furthermore, since the THESEUS satellite can localize a source within five arcminutes of its central field of view, we recorded only 1/3 of the total number of combined events in the realistic case [15,37]. We found a rate of 10 combined events per year.

3. Analysis and Results

We analyzed each mock catalog using an MCMC algorithm. We considered both events with a detected electromagnetic counterpart (bright sirens) and those without the direct redshift information (dark sirens). When we knew the redshift from the detection of GRB, the single event likelihood was [9,38]
p ( d i | Θ ) = p ( d i | D L ) p pop ( D L | z , Θ ) p ( z , z i ) dzd D L p det ( D L ) p pop ( D L | z , Θ ) p ( z , z i ) dzd D L ,
where p ( z , z i ) = δ ( z z i ) with zi being the redshift associated with the GRB. Θ is the set of cosmological parameters, and p pop ( D L | z , Θ ) = δ ( D L d L th ( z , Θ ) ) . Furthermore, the denominator is a normalization factor that takes into account the selection effects [38]. To study the dark sirens case, we assumed to know prior redshift information related to the distribution p(z), and then we marginalized over this distribution [8,9]. In this case, the likelihood is
p ( d i |   Θ   ) = 0 z max p ( d i | d L th ( z i , Θ ) ) p obs ( z i | Θ ) dz i
where the probability of the prior distribution of the redshift, p o b s ( z i | Θ ) , is obtained from the observed events and already includes detector selection effects [8].
In Table 1, we report the results obtained after ten observation years for all the models considered and for the bright and dark sirens, respectively.
It is worth stressing that we always recovered our fiducial cosmological model within the 68% confidence interval. Independently of the model used in the statistical analysis, we obtained an accuracy of ∼1% with bright sirens and reached ∼0.1% with dark sirens. This accuracy will be competitive with respect to the other cosmological probes to solve the Hubble tension [39]. However, when we consider the constraints on the additional parameters, in the non-flat ω CDM and interacting DE models, the parameters ω D E and ξ will be constrained with an accuracy worse than current bounds [21,22,24]. In the case of the time-varying gravitational constant model, the bound on the parameter δ G was one order of magnitude higher than current constraints [22], whereas we showed that ET would also be able to improve the bounds in the emergent DE model. In particular, we had an improvement of a factor 46 in the additional cosmological parameter Δ with respect to the current analysis [28]. For a more detailed comparison see [30].

4. Discussion and Conclusions

We used mock catalogs of GW events from BNSs to test the capabilities of ET on constraining the ΛCDM cosmological model and provide insight into dark energy models. Namely, we investigated the non-flat ΛCDM, the non-flat ω CDM, the interacting dark energy, the emergent dark energy, and the time-varying gravitational constant models. The third generation GW detector promises to constrain the Hubble constant with sub-percent accuracy [15], offering a possible solution to the Hubble tension.
We built mock catalogs containing GW events considering one, five, and ten years of observational runs, and SNR thresholds equal to 9. Additionally, starting from each of those three mock catalogs, we extracted a mock catalog of GW events with an associated GRB detected using the THESEUS satellite.
In the analysis, we distinguished the catalogs depending on whether the redshift information comes from the GRB (bright sirens) or the BNS merger rate (dark sirens). We assumed the rate is a priori known to follow the SFR. Although, realistically, the redshift evolution of the merger rate will be uncertain, prior knowledge of the SFR from other astrophysical observations will provide valuable information for standard siren analyses.
Our results show the huge capability of ET to solve the Hubble tension independently of the theoretical framework chosen, but also point out that, to strongly constrain the DE models we have considered, ET will need to be complemented with other datasets. The ET standard sirens will represent an alternative approach to constrain the cosmological parameters and the DE models; moreover, they will be affected by different systematics compared to the analyses based on classical electromagnetic standard candles.

Author Contributions

Methodology, M.C. and I.d.M.; writing—original draft preparation, M.C.; writing—review and editing, I.d.M. and D.V.; supervision, I.d.M., D.V. and S.C.; All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

M.C., D.V., and S.C. acknowledge the support of Istituto Nazionale di Fisica Nu-24 cleare (INFN) iniziative specifiche MOONLIGHT2, QGSKY, and TEONGRAV. I.D.M. acknowledges 25 support from Ayuda IJCI2018-036198-I funded by MCIN/AEI/10.13039/501100011033 and 26 FSE “ElFSE invierte en tu futuro” o financiado por la Unión Europea “NextGen- 27 erationEU”/PRTR. IDM is also supported by the project PID2021-122938NB-I00 funded by the Span- 28 ish “Ministerio de Ciencia e Innovación” and FEDER “A way of making Europe”, and by the project 29 SA096P20 Junta de Castilla y León. D.V. also acknowledges the FCT project with ref. number 30 PTDC/FIS-AST/0054/2021.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Abernathy, M.R.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R.X.; et al. GW150914: The Advanced LIGO Detectors in the Era of First Discoveries. Phys. Rev. Lett. 2016, 116, 131103. [Google Scholar] [CrossRef] [Green Version]
  2. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Abernathy, M.R.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R.X.; et al. Observation of Gravitational Waves from a Binary Black Hole Merger. Phys. Rev. Lett. 2016, 116, 061102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R.X.; Adya, V.B.; et al. Gravitational Waves and Gamma-Rays from a Binary Neutron Star Merger: GW170817 and GRB 170817A. Astrophys. J. Lett. 2017, 848, L13. [Google Scholar] [CrossRef] [Green Version]
  4. Schutz, B.F. Determining the Hubble Constant from Gravitational Wave Observations. Nature 1986, 323, 310–311. [Google Scholar] [CrossRef] [Green Version]
  5. Holz, D.; Hughes, S.A. Using Gravitational-Wave Standard Sirens. Astrophys. J. 2005, 629, 15–22. [Google Scholar] [CrossRef]
  6. Weinberg, S. Cosmology; Oxford University Press: Oxford, UK, 2008. [Google Scholar]
  7. Abbott, R.; Abe, H.; Acernese, F.; Ackley, K.; Adhikari, N.; Adhikari, R.X.; Adkins, V.K.; Adya, V.B.; Affeldt, C.; Agarwal, D.; et al. Constraints on the Cosmic Expansion History from GWTC-3. arXiv 2021, arXiv:2111.03604. [Google Scholar]
  8. Ding, X.; Biesiada, M.; Zheng, X.; Liao, K.; Li, Z.; Zhu, Z.-H. Cosmological Inference from Standard Sirens without Redshift Measurements. J. Cosmol. Astropart. Phys. 2019, 2019, 033. [Google Scholar] [CrossRef] [Green Version]
  9. Ye, C.; Fishbach, M. Cosmology with Standard Sirens at Cosmic Noon. Phys. Rev. D 2021, 104, 043507. [Google Scholar] [CrossRef]
  10. Verde, L.; Treu, T.; Riess, A.G. Tensions between the Early and Late Universe. Nat. Astron. 2019, 3, 891–895. [Google Scholar] [CrossRef] [Green Version]
  11. Riess, A.G.; Casertano, S.; Yuan, W.; Bowers, J.B.; Macri, L.; Zinn, J.C.; Scolnic, D. Cosmic Distances Calibrated to 1% Precision with Gaia EDR3 Parallaxes and Hubble Space Telescope Photometry of 75 Milky Way Cepheids Confirm Tension with ΛCDM. Astrophys. J. 2021, 908, L6. [Google Scholar] [CrossRef]
  12. Di Valentino, E.; Mena, O.; Pan, S.; Visinelli, L.; Yang, W.; Melchiorri, A.; Mota, D.F.; Riess, A.G.; Silk, J. In the Realm of the Hubble Tension-a Review of Solutions. Class Quantum Gravity 2021, 38, 153001. [Google Scholar] [CrossRef]
  13. Planck Collaboration; Aghanim, N.; Akrami, Y.; Ashdown, M.; Aumont, J.; Baccigalupi, C.; Ballardini, M.; Banday, A.J.; Barreiro, R.B.; Bartolo, N.; et al. Planck 2018 Results. VI. Cosmological Parameters. Astron. Astrophys. 2020, 641, A6. [Google Scholar] [CrossRef] [Green Version]
  14. Maggiore, M.; Broeck, C.V.D.; Bartolo, N.; Belgacem, E.; Bertacca, D.; Bizouard, M.A.; Branchesi, M.; Clesse, S.; Foffa, S.; García-Bellido, J.; et al. Science Case for the Einstein Telescope. J. Cosmol. Astropart. Phys. 2020, 2020, 050. [Google Scholar] [CrossRef] [Green Version]
  15. Stratta, G.; Ciolfi, R.; Amati, L.; Bozzo, E.; Ghirlanda, G.; Maiorano, E.; Nicastro, L.; Rossi, A.; Vinciguerra, S.; Frontera, F.; et al. THESEUS: A Key Space Mission Concept for Multi-Messenger Astrophysics. Adv. Space Res. 2018, 62, 662–682. [Google Scholar] [CrossRef] [Green Version]
  16. Amati, L.; O’Brien, P.; Götz, D.; Bozzo, E.; Santangelo, A.; Tanvir, N.; Frontera, F.; Mereghetti, S.; Osborne, J.P.; Blain, A.; et al. The THESEUS Space Mission: Science Goals, Requirements and Mission Concept. Exp. Astron. 2021, 52, 183–218. [Google Scholar] [CrossRef]
  17. Stratta, G.; Amati, L.; Branchesi, M.; Ciolfi, R.; Tanvir, N.; Bozzo, E.; Götz, D.; O’Brien, P.; Santangelo, A. Breakthrough Multi-Messenger Astrophysics with the THESEUS Space Mission. Galaxies 2022, 10, 60. [Google Scholar] [CrossRef]
  18. Abdalla, E.; Abellán, G.F.; Aboubrahim, A.; Agnello, A.; Akarsu, Ö.; Akrami, Y.; Alestas, G.; Aloni, D.; Amendola, L.; Anchordoqui, L.A.; et al. Cosmology Intertwined: A Review of the Particle Physics, Astrophysics, and Cosmology Associated with the Cosmological Tensions and Anomalies. J. High Energy Astrophys. 2022, 34, 49–211. [Google Scholar] [CrossRef]
  19. Di Valentino, E.; Melchiorri, A.; Silk, J. Investigating Cosmic Discordance. Astrophys. J. 2021, 908, L9. [Google Scholar] [CrossRef]
  20. Copeland, E.J.; Sami, M.; Tsujikawa, S. Dynamics of Dark Energy. Int. J. Mod. Phys. D 2006, 15, 1753–1935. [Google Scholar] [CrossRef] [Green Version]
  21. Pan, S.; Yang, W.; Di Valentino, E.; Saridakis, E.N.; Chakraborty, S. Interacting Scenarios with Dynamical Dark Energy: Observational Constraints and Alleviation of the H0 Tension. Phys. Rev. D 2019, 100, 103520. [Google Scholar] [CrossRef] [Green Version]
  22. Gao, L.-Y.; Zhao, Z.-W.; Xue, S.-S.; Zhang, X. Relieving the H_0 Tension with a New Interacting Dark Energy Model. J. Cosmol. Astropart. Phys. 2021, 2021, 005. [Google Scholar] [CrossRef]
  23. Valiviita, J.; Majerotto, E.; Maartens, R. Large-Scale Instability in Interacting Dark Energy and Dark Matter Fluids. J. Cosmol. Astropart. Phys. 2008, 2008, 020. [Google Scholar] [CrossRef] [Green Version]
  24. Di Valentino, E.; Melchiorri, A.; Mena, O.; Vagnozzi, S. Interacting Dark Energy in the Early 2020s: A Promising Solution to the H_0 and Cosmic Shear Tensions. Phys. Dark Universe 2020, 30, 100666. [Google Scholar] [CrossRef]
  25. Jin, S.-J.; Zhu, R.-Q.; Wang, L.-F.; Li, H.-L.; Zhang, J.-F.; Zhang, X. Impacts of Gravitational-Wave Standard Siren Observations from Einstein Telescope and Cosmic Explorer on Weighing Neutrinos in Interacting Dark Energy Models. Commun. Theor. Phys. 2022, 74, 105404. [Google Scholar] [CrossRef]
  26. Weinberg, S. Asymptotically Safe Inflation. Phys. Rev. D 2010, 81, 083535. [Google Scholar] [CrossRef] [Green Version]
  27. Li, X.; Shafieloo, A. Evidence for Emergent Dark Energy. Astrophys. J. 2020, 902, 58. [Google Scholar] [CrossRef]
  28. Li, X.; Shafieloo, A. A Simple Phenomenological Emergent Dark Energy Model Can Resolve the Hubble Tension. Astrophys. J. 2019, 883, L3. [Google Scholar] [CrossRef] [Green Version]
  29. Yang, W.; Di Valentino, E.; Pan, S.; Shafieloo, A.; Li, X. Generalized Emergent Dark Energy Model and the Hubble Constant Tension. Phys. Rev. D 2021, 104, 063521. [Google Scholar] [CrossRef]
  30. Califano, M.; de Martino, I.; Vernieri, D.; Capozziello, S. Exploiting the Einstein Telescope to Solve the Hubble Tension. arXiv 2022, arXiv:2208.13999. [Google Scholar]
  31. Califano, M.; de Martino, I.; Vernieri, D.; Capozziello, S. Constraining ΛCDM Cosmological Parameters with Einstein Telescope Mock Data. Mon. Not. R. Astron. Soc. 2022, 518, 3372–3385. [Google Scholar] [CrossRef]
  32. Vangioni, E.; Olive, K.A.; Prestegard, T.; Silk, J.; Petitjean, P.; Mandic, V. The Impact of Star Formation and Gamma-Ray Burst Rates at High Redshift on Cosmic Chemical Evolution and Reionization. Mon. Not. R. Astron. Soc. 2015, 447, 2575–2587. [Google Scholar] [CrossRef] [Green Version]
  33. O’Shaughnessy, R.; Belczynski, K.; Kalogera, V. Short Gamma-Ray Bursts and Binary Mergers in Spiral and Elliptical Galaxies: Redshift Distribution and Hosts. Astrophys. J. 2008, 675, 566–585. [Google Scholar] [CrossRef] [Green Version]
  34. Dalal, N.; Holz, D.E.; Hughes, S.A.; Jain, B. Short GRB and Binary Black Hole Standard Sirens as a Probe of Dark Energy. Phys. Rev. D 2006, 74, 063006. [Google Scholar] [CrossRef] [Green Version]
  35. Speri, L.; Tamanini, N.; Caldwell, R.R.; Gair, J.R.; Wang, B. Testing the Quasar Hubble Diagram with LISA Standard Sirens. Phys. Rev. D 2021, 103, 083526. [Google Scholar] [CrossRef]
  36. Hjorth, J.; Levan, A.J.; Tanvir, N.R.; Lyman, J.D.; Wojtak, R.; Schrøder, S.L.; Mandel, I.; Gall, C.; Bruun, S.H. The Distance to NGC 4993: The Host Galaxy of the Gravitational-wave Event GW170817. Astrophys. J. 2017, 848, L31. [Google Scholar] [CrossRef] [Green Version]
  37. Belgacem, E.; Dirian, Y.; Foffa, S.; Howell, E.J.; Maggiore, M.; Regimbau, T. Cosmology and Dark Energy from Joint Gravitational Wave-GRB Observations. J. Cosmol. Astropart. Phys. 2019, 2019, 015. [Google Scholar] [CrossRef] [Green Version]
  38. Mandel, I.; Farr, W.M.; Gair, J.R. Extracting Distribution Parameters from Multiple Uncertain Observations with Selection Biases. Mon. Not. R. Astron. Soc. 2019, 486, 1086–1093. [Google Scholar] [CrossRef] [Green Version]
  39. Moresco, M.; Amati, L.; Amendola, L.; Birrer, S.; Blakeslee, J.P.; Cantiello, M.; Cimatti, A.; Darling, J.; Della Valle, M.; Fishbach, M.; et al. Unveiling the Universe with emerging cosmological probes. Living Rev. Relativ. 2022, 25, 6. [Google Scholar] [CrossRef]
Table 1. The median value and the 68% confidence level of the posterior distributions of the parameters of our models for SNR equal to 9 and ten years of observations, as obtained from the MCMC analyses carried out on mock catalog collecting the bright and dark sirens, respectively.
Table 1. The median value and the 68% confidence level of the posterior distributions of the parameters of our models for SNR equal to 9 and ten years of observations, as obtained from the MCMC analyses carried out on mock catalog collecting the bright and dark sirens, respectively.
Non-Flat Λ CDM
H 0 Ω k , 0 Ω Λ , 0 -
Bright Sirens 67.49 0.87 + 0.70 0.11 0.15 + 0.16 0.74 0.15 + 0.12 -
Dark Sirens 67.68 0.03 + 0.04 0.00 0.01 + 0.01 0.69 0.01 + 0.01 -
Non-Flat ω CDM
H 0 Ω k , 0 Ω Λ , 0 ω D E
Bright Sirens 67.49 0.87 + 0.70 0.05 0.17 + 0.19 0.66 0.16 + 0.20 1.35 0.98 + 0.84
Dark Sirens 67.68 0.05 + 0.06 0.01 0.02 + 0.02 0.68 0.03 + 0.03 0.95 0.11 + 0.09
Interacting Dark Energy
H 0 Ω m , 0 ξ -
Bright Sirens 67.55 1.03 + 1.02 0.24 0.14 + 0.13 0.76 0.92 + 0.83 -
Dark Sirens 67.70 0.05 + 0.05 0.32 0.01 + 0.01 0.02 0.06 + 0.06 -
Time-Varying Gravitational Constant
H 0 Ω m , 0 δ G -
Bright Sirens 67.81 0.93 + 0.97 0.29 0.07 + 0.10 0.26 0.46 + 0.42 -
Dark Sirens 67.65 0.04 + 0.04 0.31 0.01 + 0.01 0.02 0.02 + 0.02 -
Emergent Dark Energy
H 0 Ω m , 0 Δ -
Bright Sirens 67.51 0.92 + 0.81 0.36 0.06 + 0.05 0.21 0.83 + 0.89 -
Dark Sirens 67.66 0.03 + 0.03 0.310 0.002 + 0.002 0.00 0.01 + 0.01 -
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Califano, M.; de Martino, I.; Vernieri, D.; Capozziello, S. Forecasts for ΛCDM and Dark Energy Models through Einstein Telescope Standard Sirens. Phys. Sci. Forum 2023, 7, 20. https://doi.org/10.3390/ECU2023-14032

AMA Style

Califano M, de Martino I, Vernieri D, Capozziello S. Forecasts for ΛCDM and Dark Energy Models through Einstein Telescope Standard Sirens. Physical Sciences Forum. 2023; 7(1):20. https://doi.org/10.3390/ECU2023-14032

Chicago/Turabian Style

Califano, Matteo, Ivan de Martino, Daniele Vernieri, and Salvatore Capozziello. 2023. "Forecasts for ΛCDM and Dark Energy Models through Einstein Telescope Standard Sirens" Physical Sciences Forum 7, no. 1: 20. https://doi.org/10.3390/ECU2023-14032

Article Metrics

Back to TopTop