Next Article in Journal
Phase Transitions in the Interacting Relativistic Boson Systems
Next Article in Special Issue
Spherically Symmetric C3 Matching in General Relativity
Previous Article in Journal
Universality in the Exact Renormalization Group: Comparison to Perturbation Theory
Previous Article in Special Issue
Solving the Mystery of Fast Radio Bursts: A Detective’s Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Binary Neutron-Star Mergers with a Crossover Transition to Quark Matter

1
Department of Physics and Astronomy, Center for Astrophysics, University of Notre Dame, Notre Dame, IN 46556, USA
2
Center for Quantum Spacetime, Sogang University, Seoul 04107, Republic of Korea
3
Center for Research Computing, University of Notre Dame, Notre Dame, IN 46556, USA
4
National Center for Computational Sciences, Oak Ridge National Laboratory, Oak Ridge, TN 37830, USA
*
Author to whom correspondence should be addressed.
Universe 2023, 9(9), 410; https://doi.org/10.3390/universe9090410
Submission received: 7 July 2023 / Revised: 1 August 2023 / Accepted: 30 August 2023 / Published: 7 September 2023
(This article belongs to the Special Issue Remo Ruffini Festschrift)

Abstract

:
This paper summarizes recent work on the possible gravitational-wave signal from binary neutron-star mergers in which there is a crossover transition to quark matter. Although this is a small piece of a much more complicated problem, we discuss how the power spectral density function may reveal the presence of a crossover transition to quark matter.

1. Introduction

I am honored to have a chance to contribute to this Festschrift in honor of Remo Ruffini’s 80th birthday. I have enjoyed collaboration with Remo on papers [1,2] exploring the physics of the X-ray afterglow associated gamma-ray bursts. Indeed, this collaboration inspired me to return to simulations of the relativistic hydrodynamics associated with binary neutron stars and their merger. As Remo has correctly pointed out during this Festschrift and elsewhere, the electromagnetic evolution will dominate the dynamics of binary neutron-star mergers, and moreover, there are enormous uncertainties associated with detecting and calculating the gravitational radiation emanating from binary neutron-star mergers. Nevertheless, in this presentation, I describe recent work [3] with my collaborators in which we have simulated the relativistic merger of neutron stars and explored effects on the emergent gravitational waves of a crossover transition to quark matter.
It has been discussed for some time that neutron stars (NSs) within binary systems could be used to probe the equation of state (EoS) at high densities (e.g., Refs. [4,5]). Gravitational waves (GWs) from the GW170817 event by the LIGO-Virgo Collaboration [6,7] may have provided new insight into neutron-star matter [8]. Also, NS masses and radii determined by the NICER mission constrain the EoS of nuclear matter [9,10,11].
Moreover, differences in the EoS can lead to a variety of observable effects (cf. [12]). Such changes in the EoS may lead to a change in the maximum peak frequency f p e a k (sometimes denoted as f 2 ) in the inferred power spectral density (PSD) [13,14,15]. A shift may violate the proposed universality relations between f p e a k and tidal deformability for pure hadronic EoSs [16,17,18,19,20,21]. In [3], we analyzed how such an observed shift might also probe the quark matter phase. This is, however, model dependent (e.g., [22,23]) and depends somewhat on the time duration of the merged system [12,24,25].
There have been many recent works considering EoS effects on the detected GWs. Some have considered the formation of quark matter [12,13,14,23,24,25,26,27,28,29,30,31,32,33]. Often these studies, however, were limited to a first-order phase transition. In a first order transition, a mixed phase of quarks and hadrons develops. This mixed phase diminishes the pressure support of the remnant, resulting in a prompt collapse. However, a crossover or a weak first-order transition remains a possibility [34,35,36,37,38]. The matter pressure during the crossover could be enormous. This could extend the duration of the postmerger phase. We proposed that observing such a long-duration post-merger system might signal both the phase-transition order and the strength of the coupling of quark matter [3].
In Ref. [3], we calculated the GW signal from the postmerger phase and showed that it is sensitive to the presence of quark matter in the equation of state. We demonstrated that the properties of quark matter in the crossover phase increase the duration of the postmerger GW emission. Hence, this probes the properties of quark matter. Various parameterizations of the quark-hadron crossover (QHC19) EoS of [39] were investigated in Ref. [3]. A similar study was made [40] based on the more recent version of the (QHC21) EoS with similar conclusions. The crossover is treated as continuous in the QHC19 EoS. In Ref. [3], the maximum chirp frequency f m a x , the tidal deformability, and the peak in the power spectral density f p e a k were used to identify observational characteristics of a crossover to quark matter during mergers of equal-mass binary neutron stars. The crucial high frequency range (1–4 kHz) is associated with the postmerger gravitational waves. Although this frequency is not within the sensitivity limits of the LIGO/aVirgo/KAGRA observatories, the next generation of gravitational-wave detectors, e.g., the Einstein Telescope [41] and Cosmic Explorer [42], should be sensitive to such high frequency emissions. We argue that this observation in the next generation detectors might indicate both the order of the transition and the parameters characterizing the crossover to quark matter.

2. Equations of State

At high baryon density a non-perturbative approach to QCD is required. This approach must include chiral symmetry breaking [43], the generation of constituent quark masses, quark pairing, the possibility of color superconductivity [44], etc. In the hadronic regime, we considered both the SLy [45] and the GNH3 [46] EoSs. These bracket the properties of an extremely soft or a rather stiff equation of state.
The QHC19 EoS is based upon the the NJL Lagrangian [47,48,49]. The four coupling constants are: (1) (G) the scalar coupling; (2) (K) the coefficient of the Kobayashi-Maskawa-’t Hooft vertex; (3) ( g v ) the vector coupling for quark repulsion; and (4) (H) the di-quark strength. Only two coupling constants ( g v / G and H / G ), were used to construct various versions of the EoS. The three parameter sets used in Ref. [3] are labeled as [39]: QHC19B [ ( g V , H ) = ( 0.8 , 1.49 ) ], QHC19C [ ( g V , H ) = ( 1.0 , 1.55 ) ], and QHC19D [ ( g V , H ) = ( 1.2 , 1.61 ) ]. At the crossover densities ( 2 n 0 < n < 5 n 0 ), the pressure is given by fifth-order polynomials in terms of the baryonic chemical potential.

3. Simulation Details

Binary merger simulations were run in [3] using the Einstein Toolkit [50] numerical relativity software. This includes full general relativity in three spatial dimensions with differential equations based upon the BSSN-NOK framework [51,52,53,54,55]. The hydrodynamics was evolved with the use of the GRHydro code [56,57,58] based on the Valencia formulation [59,60]. The initial conditions were generated using LORENE [61,62]. The thorn Carpet [63,64] was used for adaptive mesh refinement based upon six mesh refinement levels and a minimum grid of 0.3125 in Cactus units (≈ 461 m ). A constant adiabatic index Γ th = 1.8 was used to account for the thermal pressure in GRHydro as described in Ref. [65].
The Newman–Penrose formalism was employed to extract the gravitational waves emitted during the binary merger. This minimizes numerical noise by fitting a multipole expansion in spherical harmonics of the Weyl scalar Ψ 4 ( l , m ) ( θ , ϕ , t ) = h ¨ + ( l , m ) ( θ , ϕ , t ) + i h ¨ × ( l , m ) ( θ , ϕ , t ) . The two polarizations of the strain h + ( θ , ϕ , t ) and h × ( θ , ϕ , t ) result from a sum over the ( l , m ) modes followed by integrating twice. The neutron star models were based upon baryonic masses of M B = 1.45 , 1.50 , 1.55 M . These were chosen because gravitational masses associated with these baryonic masses for various equations of state are similar 1.35 1.4 M . Simulations began at an initial coordinate separation of 45 km between centers.
In Ref. [3] it was shown that even during inspiral, the central densities in the neutron stars achieved densities in the crossover range (2–5 n 0 ). During the merger, the maximum density increases until it exceeds ∼5–6  n 0 . The central region of the system then collapses.
It was also shown in Ref. [3] that the postmerger GW emission continues for a much longer time for the simulations with a QHC EoS. When going from QHCB to QHCC the postmerger GW emission becomes longer, corresponding to increasing the quark coupling. This led to the suggestion that the strength of the quark–matter couplings might be deduced from the duration of the post merger phase. Indeed, the lifetime of the postmerger intermediate hyper-massive neutron star (HMNS) depends rather significantly on the stiffness of the equation of state at the crossover densities. One interesting finding is that the postmerger remnants from mergers including the QHC19D EoS had so much pressure that no black hole formed during the simulations. As the EoS stiffness within the QHC models increased lifetimes of their HMNS remnants were apparent. Even the QHC19B EoS produces a much longer postmerger duration than that of a pure hadronic EoSs.
A waveform analysis of the strain can be performed in the frequency domain. This highlights the dominant frequencies of the waveform. Specifically, the effective Fourier amplitude is obtained from
h ˜ + , × ( f ) = h + , × ( t ) e i 2 π f t d t .
This is presented in Figure 1, which shows an example of the normalized power spectral density 2 h ˜ ( f ) f 1 / 2 [66] based upon the simulations of Ref. [3]. The lower blue and orange curves show the anticipated sensitivity of the Einstein Telescope and Cosmic Explorer, respectively, while the upper green curve shows the current LIGO sensitivity. The initial inspiral up to contact between the merging neutron stars ends with the first peak at around 1 kHz. For probing the crossover to quark matter, however, the peaks, f p e a k , at around 2.5–3.5 kHz are most useful. These arise from the extended postmerger phase. The amplitude of f p e a k correlates with the time duration of the postmerger remnant. Therefore, it correlates with the strength of the coupling constants in the QHC19 equations of state. As discussed in [3], one can also infer the maximum chirp strain amplitude, f m a x = 1 2 π d ϕ d t | m a x , where ϕ is the phase of the strain (see [66]). This is not apparent in the PSD, but is deduced from the phase of the strain during the merger. Although this is referred to at the maximum chirp strain, this is not to be confused with the instantaneous gravitational-wave frequency at the time of merger.

4. Discussion and Conclusions

Although the amplitude of the f p e a k PSD becomes larger for crossover equations of state with increasing coupling strengths, this might also be realized in other equations of state as demonstrated in [66]. Therefore, one desires another signature to uniquely show the formation of quark matter. In [3], it was pointed out that the QHC19 equations of state show behavior consistent with a soft EoS at low density, 3 n 0 . This affects the merger regime of f m a x . On the other hand, the postmerger phase represented in the f p e a k frequency exhibits the behavior of a stiff EoS.
This dual nature of the QHC19 EoSs might be revealed by correlating f m a x and f p e a k in a GW event [3]. This is illustrated in Figure 2. For pure hadronic EoSs, there appears to be a linear correlation between f m a x and f p e a k . However, a crossover EoS deviates from this correlation as indicated by the circled points on this figure. Thus, an observation of events in the circled region might indicate the crossover to quark matter. We note, however, that this deviation is not entirely robust as an indicator. For example, the hadronic Sly EoS also deviates from the linear relation. What is needed is a more exhaustive set of calculations to better clarify this trend. That, however, is left to a future work.
Additionally, in Ref. [3], the relation between f p e a k and the pseudo-averaged rest-mass density [17,66] was considered. For this case, the f p e a k frequencies tend to cluster in a region in between a soft and stiff EoS [3]. Hence, although there are enormous uncertainties in this suggestion, observing a transition from soft to stiffness in the correlations of f m a x and f p e a k could indicate that quark matter had formed during the merger. Moreover, the amplitude of the PSD at the frequency of f p e a k may suggest the quark–matter coupling strengths.

Author Contributions

Conceptualization, H.I.K., I.-S.S., A.K. and G.J.M. methodology, H.I.K. and A.K.; software, H.I.K., I.-S.S. and A.K.; validation, A.K., I.-S.S. and H.I.K.; formal analysis, H.I.K., A.K., G.J.M. and I.-S.S.; investigation, A.K., H.I.K., I.-S.S. and G.J.M.; resources, H.I.K., I.-S.S., G.J.M.; data curation, A.K., H.I.K. and I.-S.S. writing—original draft preparation, G.J.M.; writing—review and editing, G.J.M., I.-S.S. and A.K.; visualization, A.K. and H.I.K.; supervision, G.J.M. and H.I.K.; project administration, G.J.M. and H.I.K.; funding acquisition, G.J.M. and H.I.K. All authors have read and agreed to the published version of the manuscript.

Funding

Work at the Center for Astrophysics of the University of Notre Dame is supported by the U.S. Department of Energy under Nuclear Theory Grant No. DE-FG02-95-ER40934. This research was supported in part by the Notre Dame Center for Research Computing through the high performance computing resources. The work of H.I.K. was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education through the Center for Quantum Spacetime (CQUeST) of Sogang University (Grant No. NRF-2020R1A6A1A03047877). This research used resources of the Oak Ridge Leadership Computing Facility at the Oak Ridge National Laboratory, which is supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC05-00OR22725.

Data Availability Statement

The DOE will provide public access to these results in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan, accessed on 28 August 2023).

Acknowledgments

H.I.K. graciously thanks Jinho Kim and Chunglee Kim for continuous support. This manuscript has in part been authored by UT-Battelle, LLC under Contract No. DE-AC05-00OR22725 with the U.S. Department of Energy. The publisher, by accepting the article for publication, acknowledges that the U.S. Government retains a non-exclusive, paid up, irrevocable, world-wide license to publish or reproduce the published form of the manuscript, or allow others to do so, for U.S. Government purposes. Software used was as follows The Einstein Toolkit (Ref. [50]; https://einsteintoolkit.org, accessed on 28 August 2023), LORENE (Refs. [61,62]), PyCactus (https://bitbucket.org/GravityPR/pycactus, accessed on 28 August 2023), and TOVsolver (https://github.com/amotornenko/TOVsolver, accessed on 28 August 2023).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ruffini, R.; Karlica, M.; Sahakyan, N.; Rueda, J.A.; Wang, Y.; Mathews, G.J.; Bianco, C.L.; Muccino, M. A GRB Afterglow Model Consistent with Hypernova Observations. Astrophys. J. 2018, 869, 101. [Google Scholar] [CrossRef]
  2. Ruffini, R.; Moradi, R.; Rueda, J.A.; Li, L.; Sahakyan, N.; Chen, Y.C.; Wang, Y.; Aimuratov, Y.; Becerra, L.; Bianco, C.L.; et al. The morphology of the X-ray afterglows and of the jetted GeV emission in long GRBs. MNRAS 2021, 504, 5301–5326. [Google Scholar] [CrossRef]
  3. Kedia, A.; Kim, H.I.; Suh, I.S.; Mathews, G.J. Binary neutron star mergers as a probe of quark-hadron crossover equations of state. Phys. Rev. D 2022, 106, 103027. [Google Scholar] [CrossRef]
  4. Baiotti, L. Gravitational waves from neutron star mergers and their relation to the nuclear equation of state. Prog. Part. Nucl. Phys. 2019, 109, 103714. [Google Scholar] [CrossRef]
  5. Radice, D.; Bernuzzi, S.; Perego, A. The Dynamics of Binary Neutron Star Mergers and GW170817. Annu. Rev. Nucl. Part. Sci. 2020, 70, 95–119. [Google Scholar] [CrossRef]
  6. 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. GW170817: Observation of Gravitational Waves from a Binary Neutron Star Inspiral. Phys. Rev. Lett. 2017, 119, 161101. [Google Scholar] [CrossRef]
  7. 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. GW170817: Measurements of Neutron Star Radii and Equation of State. Phys. Rev. Lett. 2018, 121, 161101. [Google Scholar] [CrossRef] [PubMed]
  8. Lattimer, J.M. The Nuclear Equation of State and Neutron Star Masses. Annu. Rev. Nucl. Part. Sci. 2012, 62, 485–515. [Google Scholar] [CrossRef]
  9. Bogdanov, S.; Guillot, S.; Ray, P.S.; Wolff, M.T.; Chakrabarty, D.; Ho, W.C.G.; Kerr, M.; Lamb, F.K.; Lommen, A.; Ludlam, R.M.; et al. Constraining the Neutron Star Mass-Radius Relation and Dense Matter Equation of State with NICER. I. The Millisecond Pulsar X-ray Data Set. Astrophys. J. Lett. 2019, 887, L25. [Google Scholar] [CrossRef]
  10. Bogdanov, S.; Dittmann, A.J.; Ho, W.C.G.; Lamb, F.K.; Mahmoodifar, S.; Miller, M.C.; Morsink, S.M.; Riley, T.E.; Strohmayer, T.E.; Watts, A.L.; et al. Constraining the Neutron Star Mass-Radius Relation and Dense Matter Equation of State with NICER. III. Model Description and Verification of Parameter Estimation Codes. Astrophys. J. Lett. 2021, 914, L15. [Google Scholar] [CrossRef]
  11. Miller, M.C.; Lamb, F.K.; Dittmann, A.J.; Bogdanov, S.; Arzoumanian, Z.; Gendreau, K.C.; Guillot, S.; Ho, W.C.G.; Lattimer, J.M.; Loewenstein, M.; et al. The Radius of PSR J0740+6620 from NICER and XMM-Newton Data. Astrophys. J. Lett. 2021, 918, L28. [Google Scholar] [CrossRef]
  12. Weih, L.R.; Hanauske, M.; Rezzolla, L. Postmerger Gravitational-Wave Signatures of Phase Transitions in Binary Mergers. Phys. Rev. Lett. 2020, 124, 171103. [Google Scholar] [CrossRef] [PubMed]
  13. Bauswein, A.; Bastian, N.U.F.; Blaschke, D.B.; Chatziioannou, K.; Clark, J.A.; Fischer, T.; Oertel, M. Identifying a First-Order Phase Transition in Neutron-Star Mergers through Gravitational Waves. Phys. Rev. Lett. 2019, 122, 061102. [Google Scholar] [CrossRef] [PubMed]
  14. Blacker, S.; Bastian, N.U.F.; Bauswein, A.; Blaschke, D.B.; Fischer, T.; Oertel, M.; Soultanis, T.; Typel, S. Constraining the onset density of the hadron-quark phase transition with gravitational-wave observations. Phys. Rev. D 2020, 102, 123023. [Google Scholar] [CrossRef]
  15. Radice, D. General-relativistic Large-eddy Simulations of Binary Neutron Star Mergers. Astrophys. J. Lett. 2017, 838, L2. [Google Scholar] [CrossRef]
  16. Breschi, M.; Bernuzzi, S.; Zappa, F.; Agathos, M.; Perego, A.; Radice, D.; Nagar, A. Kilohertz gravitational waves from binary neutron star remnants: Time-domain model and constraints on extreme matter. Phys. Rev. D 2019, 100, 104029. [Google Scholar] [CrossRef]
  17. Bauswein, A.; Janka, H.T. Measuring Neutron-Star Properties via Gravitational Waves from Neutron-Star Mergers. Phys. Rev. Lett. 2012, 108, 011101. [Google Scholar] [CrossRef]
  18. Hotokezaka, K.; Kiuchi, K.; Kyutoku, K.; Muranushi, T.; Sekiguchi, Y.i.; Shibata, M.; Taniguchi, K. Remnant massive neutron stars of binary neutron star mergers: Evolution process and gravitational waveform. Phys. Rev. D 2013, 88, 044026. [Google Scholar] [CrossRef]
  19. Bernuzzi, S.; Nagar, A.; Balmelli, S.; Dietrich, T.; Ujevic, M. Quasiuniversal Properties of Neutron Star Mergers. Phys. Rev. Lett. 2014, 112, 201101. [Google Scholar] [CrossRef]
  20. Rezzolla, L.; Takami, K. Gravitational-wave signal from binary neutron stars: A systematic analysis of the spectral properties. Phys. Rev. D 2016, 93, 124051. [Google Scholar] [CrossRef]
  21. Zappa, F.; Bernuzzi, S.; Radice, D.; Perego, A.; Dietrich, T. Gravitational-Wave Luminosity of Binary Neutron Stars Mergers. Phys. Rev. Lett. 2018, 120, 111101. [Google Scholar] [CrossRef] [PubMed]
  22. Most, E.R.; Papenfort, L.J.; Dexheimer, V.; Hanauske, M.; Schramm, S.; Stöcker, H.; Rezzolla, L. Signatures of Quark-Hadron Phase Transitions in General-Relativistic Neutron-Star Mergers. Phys. Rev. Lett. 2019, 122, 061101. [Google Scholar] [CrossRef] [PubMed]
  23. Most, E.R.; Jens Papenfort, L.; Dexheimer, V.; Hanauske, M.; Stoecker, H.; Rezzolla, L. On the deconfinement phase transition in neutron-star mergers. Eur. Phys. J. A 2020, 56, 59. [Google Scholar] [CrossRef]
  24. Liebling, S.L.; Palenzuela, C.; Lehner, L. Effects of high density phase transitions on neutron star dynamics. Class. Quantum Gravity 2021, 38, 115007. [Google Scholar] [CrossRef]
  25. Prakash, A.; Radice, D.; Logoteta, D.; Perego, A.; Nedora, V.; Bombaci, I.; Kashyap, R.; Bernuzzi, S.; Endrizzi, A. Signatures of deconfined quark phases in binary neutron star mergers. Phys. Rev. D 2021, 104, 083029. [Google Scholar] [CrossRef]
  26. Sebastian Blacker and Hristijan Kochankovski and Andreas Bauswein and Angels Ramos and Laura Tolos Thermal behavior as indicator for hyperons in binary neutron star merger remnants. arXiv 2023, arXiv:2307.03710.
  27. Mallick, R. Gravitational wave signatures of phase transition from hadronic to quark matter in isolated neutron stars and binaries. Eur. Phys. J. Web Conf. 2022, 274, 07002. [Google Scholar] [CrossRef]
  28. Ren, J.; Zhang, C. Hybrid stars may have an inverted structure. arXiv 2022, arXiv:2211.12043. [Google Scholar]
  29. Tajima, H.; Tsutsui, S.; Doi, T.M.; Iida, K. Density-Induced Hadron–Quark Crossover via the Formation of Cooper Triples. Symmetry 2023, 15, 333. [Google Scholar] [CrossRef]
  30. Kohsuke, S.; Toru, K.; Shun, F. Equation of State in Neutron Stars and Supernovae. In Handbook of Nuclear Physics; Springer: Singapore, 2023; pp. 1–51. [Google Scholar]
  31. Sebastian, B.; Andreas, B.; Stefan, T. Exploring thermal effects of the hadron-quark matter transition in neutron star mergers. arXiv 2023, arXiv:2304.01971. [Google Scholar]
  32. Bauswein, A.; Janka, H.T.; Hebeler, K.; Schwenk, A. Equation-of-state dependence of the gravitational-wave signal from the ring-down phase of neutron-star mergers. Phys. Rev. D 2012, 86, 063001. [Google Scholar] [CrossRef]
  33. Bauswein, A.; Blacker, S.; Vijayan, V.; Stergioulas, N.; Chatziioannou, K.; Clark, J.A.; Bastian, N.U.F.; Blaschke, D.B.; Cierniak, M.; Fischer, T. Equation of State Constraints from the Threshold Binary Mass for Prompt Collapse of Neutron Star Mergers. Phys. Rev. Lett. 2020, 125, 141103. [Google Scholar] [CrossRef] [PubMed]
  34. Pisarski, R.D.; Skokov, V.V. Chiral matrix model of the semi-QGP in QCD. Phys. Rev. D 2016, 94, 034015. [Google Scholar] [CrossRef]
  35. Steinheimer, J.; Schramm, S. The problem of repulsive quark interactions - Lattice versus mean field models. Phys. Lett. B 2011, 696, 257–261. [Google Scholar] [CrossRef]
  36. Hatsuda, T.; Tachibana, M.; Yamamoto, N.; Baym, G. New Critical Point Induced By the Axial Anomaly in Dense QCD. Phys. Rev. Lett. 2006, 97, 122001. [Google Scholar] [CrossRef]
  37. Aoki, Y.; Endrodi, G.; Fodor, Z.; Katz, S.D.; Szabó, K.K. The order of the quantum chromodynamics transition predicted by the standard model of particle physics. Nature 2006, 443, 675–678. [Google Scholar] [CrossRef]
  38. Baym, G.; Hatsuda, T.; Kojo, T.; Powell, P.D.; Song, Y.; Takatsuka, T. From hadrons to quarks in neutron stars: A review. Rep. Prog. Phys. 2018, 81, 056902. [Google Scholar] [CrossRef]
  39. Baym, G.; Furusawa, S.; Hatsuda, T.; Kojo, T.; Togashi, H. New Neutron Star Equation of State with Quark-Hadron Crossover. Astrophys. J. 2019, 885, 42. [Google Scholar] [CrossRef]
  40. Kojo, T.; Baym, G.; Hatsuda, T. Implications of NICER for Neutron Star Matter: The QHC21 Equation of State. Astrophys. J. 2022, 934, 46. [Google Scholar] [CrossRef]
  41. Sathyaprakash, B.; Abernathy, M.; Acernese, F.; Ajith, P.; Allen, B.; Amaro-Seoane, P.; Andersson, N.; Aoudia, S.; Arun, K.; Astone, P.; et al. Scientific objectives of Einstein Telescope. Class. Quantum Gravity 2012, 29, 124013. [Google Scholar] [CrossRef]
  42. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Abernathy, M.R.; Ackley, K.; Adams, C.; Addesso, P.; Adhikari, R.X.; Adya, V.B.; Affeldt, C.; et al. Exploring the sensitivity of next generation gravitational wave detectors. Class. Quantum Gravity 2017, 34, 044001. [Google Scholar] [CrossRef]
  43. Hatsuda, T.; Kunihiro, T. QCD phenomenology based on a chiral effective Lagrangian. Phys. Rep. 1994, 247, 221–367. [Google Scholar] [CrossRef]
  44. Alford, M.G.; Schmitt, A.; Rajagopal, K.; Schäfer, T. Color superconductivity in dense quark matter. Rev. Mod. Phys. 2008, 80, 1455–1515. [Google Scholar] [CrossRef]
  45. Chabanat, E.; Bonche, P.; Haensel, P.; Meyer, J.; Schaeffer, R. A Skyrme parametrization from subnuclear to neutron star densitiesPart II. Nuclei far from stabilities. Nucl. Phys. A 1998, 635, 231–256. [Google Scholar] [CrossRef]
  46. Glendenning, N.K. Neutron stars are giant hypernuclei? Astrophys. J. 1985, 293, 470–493. [Google Scholar] [CrossRef]
  47. Nambu, Y.; Jona-Lasinio, G. Dynamical Model of Elementary Particles Based on an Analogy with Superconductivity. I. Phys. Rev. 1961, 122, 345–358. [Google Scholar] [CrossRef]
  48. Nambu, Y.; Jona-Lasinio, G. Dynamical Model of Elementary Particles Based on an Analogy with Superconductivity. II. Phys. Rev. 1961, 124, 246–254. [Google Scholar] [CrossRef]
  49. Buballa, M. NJL-model analysis of dense quark matter [review article]. Phys. Rep. 2005, 407, 205–376. [Google Scholar] [CrossRef]
  50. Etienne, Z.; Brandt, S.R.; Diener, P.; Gabella, W.E.; Gracia-Linares, M.; Haas, R.; Kedia, A.; Alcubierre, M.; Alic, D.; Allen, G.; et al. The Einstein Toolkit. Zenodo (The "Lorentz" release, ET_2021_05) 2021. [Google Scholar] [CrossRef]
  51. Nakamura, T.; Oohara, K.; Kojima, Y. General Relativistic Collapse to Black Holes and Gravitational Waves from Black Holes. Prog. Theor. Phys. Suppl. 1987, 90, 1–218. [Google Scholar] [CrossRef]
  52. Shibata, M.; Nakamura, T. Evolution of three-dimensional gravitational waves: Harmonic slicing case. Phys. Rev. D 1995, 52, 5428–5444. [Google Scholar] [CrossRef] [PubMed]
  53. Baumgarte, T.W.; Shapiro, S.L. Numerical integration of Einstein’s field equations. Phys. Rev. D 1998, 59, 024007. [Google Scholar] [CrossRef]
  54. Alcubierre, M.; Brügmann, B.; Dramlitsch, T.; Font, J.A.; Papadopoulos, P.; Seidel, E.; Stergioulas, N.; Takahashi, R. Towards a stable numerical evolution of strongly gravitating systems in general relativity: The conformal treatments. Phys. Rev. D 2000, 62, 044034. [Google Scholar] [CrossRef]
  55. Alcubierre, M.; Brügmann, B.; Diener, P.; Koppitz, M.; Pollney, D.; Seidel, E.; Takahashi, R. Gauge conditions for long-term numerical black hole evolutions without excision. Phys. Rev. D 2003, 67, 084023. [Google Scholar] [CrossRef]
  56. Baiotti, L.; Hawke, I.; Montero, P.J.; Löffler, F.; Rezzolla, L.; Stergioulas, N.; Font, J.A.; Seidel, E. Three-dimensional relativistic simulations of rotating neutron-star collapse to a Kerr black hole. Phys. Rev. D 2005, 71, 024035. [Google Scholar] [CrossRef]
  57. Hawke, I.; Löffler, F.; Nerozzi, A. Excision methods for high resolution shock capturing schemes applied to general relativistic hydrodynamics. Phys. Rev. D 2005, 71, 104006. [Google Scholar] [CrossRef]
  58. Mösta, P.; Mundim, B.C.; Faber, J.A.; Haas, R.; Noble, S.C.; Bode, T.; Löffler, F.; Ott, C.D.; Reisswig, C.; Schnetter, E. GRHydro: A new open-source general-relativistic magnetohydrodynamics code for the Einstein toolkit. Class. Quantum Gravity 2014, 31, 015005. [Google Scholar] [CrossRef]
  59. Banyuls, F.; Font, J.A.; Ibáñez, J.M.; Martí, J.M.; Miralles, J.A. Numerical {3 + 1} General Relativistic Hydrodynamics: A Local Characteristic Approach. Astrophys. J. 1997, 476, 221–231. [Google Scholar] [CrossRef]
  60. Font, J.A. Numerical Hydrodynamics and Magnetohydrodynamics in General Relativity. Living Rev. Relativ. 2008, 11, 7. [Google Scholar] [CrossRef]
  61. Gourgoulhon, E.; Grandclément, P.; Taniguchi, K.; Marck, J.A.; Bonazzola, S. Quasiequilibrium sequences of synchronized and irrotational binary neutron stars in general relativity: Method and tests. Phys. Rev. D 2001, 63, 064029. [Google Scholar] [CrossRef]
  62. Grandclément, P.; Novak, J. Spectral Methods for Numerical Relativity. Living Rev. Relativ. 2009, 12, 1. [Google Scholar] [CrossRef] [PubMed]
  63. Schnetter, E.; Hawley, S.H.; Hawke, I. Evolutions in 3D numerical relativity using fixed mesh refinement. Class. Quantum Gravity 2004, 21, 1465–1488. [Google Scholar] [CrossRef]
  64. Schnetter, E.; Diener, P.; Dorband, E.N.; Tiglio, M. A multi-block infrastructure for three-dimensional time-dependent numerical relativity. Class. Quantum Gravity 2006, 23, S553–S578. [Google Scholar] [CrossRef]
  65. De Pietri, R.; Feo, A.; Maione, F.; Löffler, F. Modeling equal and unequal mass binary neutron star mergers using public codes. Phys. Rev. D 2016, 93, 064047. [Google Scholar] [CrossRef]
  66. Takami, K.; Rezzolla, L.; Baiotti, L. Spectral properties of the post-merger gravitational-wave signal from binary neutron stars. Phys. Rev. D 2015, 91, 064001. [Google Scholar] [CrossRef]
Figure 1. Power spectral density ( 2 h ˜ ( f ) f 1 / 2 ) vs. frequency f for various simulations. The the lower blue and orange curves show anticipated sensitivity of the Einstein Telescope and Cosmic Explorer, respectively, while the upper green curve shows the LIGO sensitivity. The first peak at around 1 kHz is the initial contact of the merging binaries. The second peaks near 2.5–3.5 kHz correspond to the long postmerger phase, f p e a k .
Figure 1. Power spectral density ( 2 h ˜ ( f ) f 1 / 2 ) vs. frequency f for various simulations. The the lower blue and orange curves show anticipated sensitivity of the Einstein Telescope and Cosmic Explorer, respectively, while the upper green curve shows the LIGO sensitivity. The first peak at around 1 kHz is the initial contact of the merging binaries. The second peaks near 2.5–3.5 kHz correspond to the long postmerger phase, f p e a k .
Universe 09 00410 g001
Figure 2. Correlation between f m a x and f p e a k . There appears to be a linear correlation for normal hadronic EoSs as indicated by the straight line. However, the existence of a crossover regime to quark matter leads to outliers from this correlation as indicated by the circled points.
Figure 2. Correlation between f m a x and f p e a k . There appears to be a linear correlation for normal hadronic EoSs as indicated by the straight line. However, the existence of a crossover regime to quark matter leads to outliers from this correlation as indicated by the circled points.
Universe 09 00410 g002
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

Mathews, G.J.; Kedia, A.; Kim, H.I.; Suh, I.-S. Binary Neutron-Star Mergers with a Crossover Transition to Quark Matter. Universe 2023, 9, 410. https://doi.org/10.3390/universe9090410

AMA Style

Mathews GJ, Kedia A, Kim HI, Suh I-S. Binary Neutron-Star Mergers with a Crossover Transition to Quark Matter. Universe. 2023; 9(9):410. https://doi.org/10.3390/universe9090410

Chicago/Turabian Style

Mathews, Grant J., Atul Kedia, Hee Il Kim, and In-Saeng Suh. 2023. "Binary Neutron-Star Mergers with a Crossover Transition to Quark Matter" Universe 9, no. 9: 410. https://doi.org/10.3390/universe9090410

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop