The quasi-Gaussian distribution of DMIGM in fast radio bursts may bias the constraints on the Hubble constant

  • Fast radio bursts (FRBs) are useful cosmological probes with numerous applications in cosmology. The distribution of the dispersion measure contribution from the intergalactic medium is a key issue. A quasi-Gaussian distribution has been used to replace the traditional Gaussian distribution, which yields promising results. However, our work suggests that there may be additional challenges in its application. Here we use 35 well-localized FRBs to constrain the Hubble constant $ H_0 $ along with two FRB-related parameters, yielding $ H_0=60.99^{+4.57}_{-4.90}\ {\rm{km\; s^{-1}}\; Mpc^{-1}} $. The best-fitting Hubble constant $ H_0 $ is smaller than the value obtained from the Cosmic Microwave Background (CMB), which may be caused by the small sample size of current FRB data. Monte Carlo simulations indicate that a set of 100 simulated FRBs provides a more precise fitting result for the Hubble constant. But the precision of the Hubble constant does not improve when further enlarging the FRB sample. Further simulations reveal a systematic deviation in the fitting results of $ H_0 $, attributed to the quasi-Gaussian distribution of dispersion measure in the intergalactic medium. Despite this, the results remain reliable within the 1σ uncertainty, assuming a sufficient number of FRB data are available.
  • 加载中
  • [1] E. Petroff, E. D. Barr, A. Jameson, E. F. Keane, M. Bailes, M. Kramer, V. Morello, D. Tabbara, W. van Straten, Publ. Astron. Soc. Austral. 33, e045 (2016) doi: 10.1017/pasa.2016.35
    [2] E. Petroff, J. W. T. Hessels, D. R. Lorimer, The Astronomy and Astrophysics Review 27(1), 4 (2019) doi: 10.1007/s00159-019-0116-6
    [3] D. Xiao, F. Wang, Z. Dai, Mechanics & Astronomy 64(4), 249501 (2021)
    [4] B. Zhang, Reviews of Modern Physics 95(3), 035005 (2024)
    [5] D. R. Lorimer, M. Bailes, M. A. McLaughlin, D. J. Narkevic, F. Crawford, Science 318(5851), 777 (2007) doi: 10.1126/science.1147532
    [6] D. Thornton, B. Stappers, M. Bailes, et al., Science 341(6141), 53 (2013) doi: 10.1126/science.1236789
    [7] D. Li, P. Wang, W. W. Zhu, et al., Nature 598(7880), 267 (2021) doi: 10.1038/s41586-021-03878-5
    [8] B. C. Andersen, K. Bandura, M. Bhardwaj, et al., The Astrophysical Journal Letters 885(1), L24 (2019) doi: 10.3847/2041-8213/ab4a80
    [9] The CHIME/FRB Collaboration, The Astrophysical Journal Supplement Series 257(2), 59 (2021) doi: 10.3847/1538-4365/ac33ab
    [10] The CHIME/FRB Collaboration, Nature 587(7832), 54 (2020) doi: 10.1038/s41586-020-2863-y
    [11] E. F. Keane, S. Johnston, S. Bhandari, et al., Nature 530(7591), 453 (2016) doi: 10.1038/nature17140
    [12] S. Chatterjee, C. J. Law, R. S. Wharton, et al., Nature 541(7635), 58 (2017) doi: 10.1038/nature20797
    [13] S. P. Tendulkar, C. G. Bassa, J. M. Cordes, et al., The Astrophysical Journal Letters 02, L7 (1211)
    [14] A. G. Riess, et al., Astrophys. J. 861(2), 126 (2018) doi: 10.3847/1538-4357/aac82e
    [15] A. G. Riess, et al., Astrophys. J. 855(2), 136 (2018) doi: 10.3847/1538-4357/aaadb7
    [16] A. G. Riess, et al., Astrophys. J. Lett. 934(1), L7 (2022) doi: 10.3847/2041-8213/ac5c5b
    [17] E. D. Valentino, O. Mena, S. Pan, et al., Classical and Quantum Gravity 38(15), 153001 (2021) doi: 10.1088/1361-6382/ac086d
    [18] J.-P. Macquart, et al., Nature 581(7809), 391 (2020) doi: 10.1038/s41586-020-2300-2
    [19] H. Gao, Z. Li, B. Zhang, The Astrophysical Journal 788(2), 189 (2014) doi: 10.1088/0004-637X/788/2/189
    [20] Zhao, et al., The Astrophysical Journal 903(2), 83 (2020) doi: 10.3847/1538-4357/abb8ce
    [21] Qiu, et al., Journal of Cosmology and Astroparticle Physics 2022(02), 006 (2022) doi: 10.1088/1475-7516/2022/02/006
    [22] Zhao, et al., First statistical measurement of the Hubble constant using unlocalized fast radio bursts, arXiv: 2212.13433[astro-ph, physics: gr-qc, physics: hep-ph] (Dec. 2022).
    [23] Zhang, et al., Mechanics & Astronomy 66(12), 120412 (2023)
    [24] M. Bhattacharya, P. Kumar, E. V. Linder, Phys. Rev. D 103, 103526 (2021) doi: 10.1103/PhysRevD.103.103526
    [25] H.-N. Lin, Y. Sang, Chin. Phys. C 45(12), 125101 (2021) doi: 10.1088/1674-1137/ac2660
    [26] Q. Wu, G. Q. Zhang, Monthly Notices of the Royal Astronomical Society: Letters 515(1), L1 (2022) doi: 10.1093/mnrasl/slac022
    [27] Z.-X. Li, H. Gao, X.-H. Ding, G.-J. Wang, B. Zhang, Nature Commun. 9(1), 3833 (2018) doi: 10.1038/s41467-018-06303-0
    [28] N. Pearson, C. Trendafilova, J. Meyers, Phys. Rev. D 103(6), 063017 (2021) doi: 10.1103/PhysRevD.103.063017
    [29] K. Masui, Nature Astronomy 7(6), 749 (2023) doi: 10.1038/s41550-023-01970-6
    [30] P. Sanghavi, other, TONE: A CHIME/FRB Outrigger Pathfinder for localizations of Fast Radio Bursts using Very Long Baseline Interferometry, arXiv: 2304.10534[astro-ph] (Apr. 2023).
    [31] Z. Li, H. Gao, J.-J. Wei, et al., The Astrophysical Journal 876(2), 146 (2019) doi: 10.3847/1538-4357/ab18fe
    [32] H.-N. Lin, X. Li, L. Tang, Chin. Phys. C 46(7), 075102 (2022) doi: 10.1088/1674-1137/ac5e92
    [33] G. Q. Zhang, H. Yu, J. H. He, F. Y. Wang, The Astrophysical Journal 900(2), 170 (2020) doi: 10.3847/1538-4357/abaa4a
    [34] J. M. Cordes, T. J. W. Lazio, NE2001.I. A New Model for the Galactic Distribution of Free Electrons and its Fluctuations.
    [35] J. X. Prochaska, Y. Zheng, Monthly Notices of the Royal Astronomical Society 485(1), 648 (2019)
    [36] F. Y. Wang, G. Q. Zhang, Z. G. Dai, Repeating fast radio burst 13(1), 4382 (2022)
    [37] W. Deng, B. Zhang, The Astrophysical Journal Letters 783(2), L35 (2014) doi: 10.1088/2041-8205/783/2/L35
    [38] R. C. Zhang, et al., Monthly Notices of the Royal Astronomical Society 501(1), 157 (2020) doi: 10.1093/mnras/staa3537
    [39] A. A. Meiksin, Rev. Mod. Phys. 81, 1405 (2009) doi: 10.1103/RevModPhys.81.1405
    [40] G. D. Becker, et al., Monthly Notices of the Royal Astronomical Society 410(2), 1096 (2010)
    [41] McQuinn, Locating the “Missing” Baryons with Extragalactic Dispersion Measure Estimates, The Astrophysical Journal 780 (2).
    [42] Y.-P. Yang, B. Zhang, Extracting host galaxy dispersion measure and constraining cosmological parameters using fast radio burst data, The Astrophysical Journal 830 (2) (2016) L31, arXiv: 1608.08154[astro-ph].
    [43] M. Jaroszynski, Fast radio bursts and cosmological tests (Apr. 2019).
    [44] Z. J. Zhang, K. Yan, C. M. Li, G. Q. Zhang, F. Y. Wang, The Astrophysical Journal 906(1), 49 (2021) doi: 10.3847/1538-4357/abceb9
    [45] H. Xu, et al., Nature 611(7937), E12 (2022) doi: 10.1038/s41586-022-05493-4
    [46] A repeating fast radio burst associated with a persistent radio source, Nature 606 (7916) (2022) 873–877.
    [47] M. Bhardwaj, et al., The Astrophysical Journal Letters 910(2), L18 (2021) doi: 10.3847/2041-8213/abeaa6
    [48] F. Kirsten, B. Marcote, K. Nimmo, et al., Nature 602(7898), 585 (2022) doi: 10.1038/s41586-021-04354-w
    [49] C. J. Law, B. J. Butler, J. X. Prochaska, et al., The Astrophysical Journal 899(2), 161 (2020) doi: 10.3847/1538-4357/aba4ac
    [50] V. Ravi, et al., Deep Synoptic Array science: a 50 Mpc fast radio burst constrains the mass of the Milky Way circumgalactic medium, arXiv: 2301.01000[astro-ph] (Jan. 2023).
    [51] Y. Li, B. Zhang, K. Nagamine, The Astrophysical Journal Letters 884(1), L26 (2019) doi: 10.3847/2041-8213/ab3e41
    [52] S. Bhandari, K. E. Heintz, K. Aggarwal, et al., The Astronomical Journal 163(2), 69 (2022) doi: 10.3847/1538-3881/ac3aec
    [53] B. Marcote, et al., Nature 577(7789), 190 (2020) doi: 10.1038/s41586-019-1866-z
    [54] K. W. Bannister, A. T. Deller, C. Phillips, et al., Science 365(6453), 565 (2019) doi: 10.1126/science.aaw5903
    [55] M. Bhardwaj, A. Y. Kirichenko, D. Michilli, The Astrophysical Journal Letters 919(2), L24 (2021) doi: 10.3847/2041-8213/ac223b
    [56] J. X. Prochaska, J.-P. Macquart, M. McQuinn, et al., Science 366(6462), 231 (2019) doi: 10.1126/science.aay0073
    [57] V. Ravi, et al., Nature 572(7769), 352 (2019) doi: 10.1038/s41586-019-1389-7
    [58] K. E. Heintz, J. X. Prochaska, S. Simha, et al., The Astrophysical Journal 903(2), 152 (2020) doi: 10.3847/1538-4357/abb6fb
    [59] W. fai Fong, Y. Dong, J. Leja, Chronicling the Host Galaxy Properties of the Remarkable Repeating FRB 919(2), L23 (2021)
    [60] L. N. Driessen, et al., Monthly Notices of the Royal Astronomical Society 527(2), 3659 (2023) doi: 10.1093/mnras/stad3329
    [61] M. Caleb, et al., Monthly Notices of the Royal Astronomical Society 524(2), 2064 (2023) doi: 10.1093/mnras/stad1839
    [62] T. Cassanelli, et al., A fast radio burst localized at detection to a galactic disk using very long baseline interferometry, arXiv: 2307.09502[astro-ph] (Jul. 2023).
    [63] I. S. Khrykin, et al., FLIMFLAM DR1: The First Constraints on the Cosmic Baryon Distribution from 8 FRB sightlines, arXiv: 2402.00505[astro-ph] (Feb. 2024).
    [64] C. J. Law, et al., Deep Synoptic Array Science: First FRB and Host Galaxy Catalog, arXiv: 2307.03344[astro-ph] (Jan. 2024).
    [65] Y.-K. Zhang, et al., FAST Observations of FRB 955(2), 142 (2023)
    [66] R. Mckinven, et al., A pulsar-like swing in the polarisation position angle of a nearby fast radio burst, arXiv: 2402.09304[astro-ph] (Feb. 2024).
    [67] A. Ferrara, S. Pandolfi, Reionization of the Intergalactic Medium, arXiv: 1409.4946[astro-ph] (Sep. 2014).
    [68] H.-N. Lin, R. Zou, Monthly Notices of the Royal Astronomical Society 520(4), 6237 (2023) doi: 10.1093/mnras/stad509
    [69] Planck 2018 results: VI. Cosmological parameters, A&A 641 (2020) A6.
    [70] D. Foreman-Mackey, D. W. Hogg, D. Lang, J. Goodman, Publications of the Astronomical Society of the Pacific 125(925), 306 (2013) doi: 10.1086/670067
    [71] L. Tang, et al., Chinese Physics C 47(8), 085105 (2023) doi: 10.1088/1674-1137/acda1c
    [72] M. Jaroszynski, Monthly Notices of the Royal Astronomical Society 484(2), 1637 (2019) doi: 10.1093/mnras/sty3529
    [73] P. Kumar, E. V. Linder, Phys. Rev. D 100, 083533 (2019) doi: 10.1103/PhysRevD.100.083533
    [74] H. Yüksel, M. D. Kistler, J. F. Beacom, A. M. Hopkins, The Astrophysical Journal 683(1), L5 (2008) doi: 10.1086/591449
  • 加载中

Figures(5) / Tables(1)

Get Citation
Gui-Yao Chen and Xin Li. The quasi-Gaussian distribution of DMIGM in fast radio bursts may bias the constraints on the Hubble constant[J]. Chinese Physics C.
Gui-Yao Chen and Xin Li. The quasi-Gaussian distribution of DMIGM in fast radio bursts may bias the constraints on the Hubble constant[J]. Chinese Physics C. shu
Milestone
Received: 2024-01-15
Article Metric

Article Views(105)
PDF Downloads(4)
Cited by(0)
Policy on re-use
To reuse of subscription content published by CPC, the users need to request permission from CPC, unless the content was published under an Open Access license which automatically permits that type of reuse.
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Email This Article

Title:
Email:

The quasi-Gaussian distribution of DMIGM in fast radio bursts may bias the constraints on the Hubble constant

  • 1. Department of Physics, Chongqing University, Chongqing 401331, China
  • 2. Chongqing Key Laboratory for Strongly Coupled Physics, Chongqing University, Chongqing 401331, China

Abstract: Fast radio bursts (FRBs) are useful cosmological probes with numerous applications in cosmology. The distribution of the dispersion measure contribution from the intergalactic medium is a key issue. A quasi-Gaussian distribution has been used to replace the traditional Gaussian distribution, which yields promising results. However, our work suggests that there may be additional challenges in its application. Here we use 35 well-localized FRBs to constrain the Hubble constant $ H_0 $ along with two FRB-related parameters, yielding $ H_0=60.99^{+4.57}_{-4.90}\ {\rm{km\; s^{-1}}\; Mpc^{-1}} $. The best-fitting Hubble constant $ H_0 $ is smaller than the value obtained from the Cosmic Microwave Background (CMB), which may be caused by the small sample size of current FRB data. Monte Carlo simulations indicate that a set of 100 simulated FRBs provides a more precise fitting result for the Hubble constant. But the precision of the Hubble constant does not improve when further enlarging the FRB sample. Further simulations reveal a systematic deviation in the fitting results of $ H_0 $, attributed to the quasi-Gaussian distribution of dispersion measure in the intergalactic medium. Despite this, the results remain reliable within the 1σ uncertainty, assuming a sufficient number of FRB data are available.

    HTML

    I.   INTRODUCTION
    • Fast radio bursts (FRBs) are bright and energetic radio pulses characterized by millisecond durations, primarily originating from extragalactic sources [14]. The first FRB signal, now named FRB 010724, was discovered by Lorimer et al. [5] in 2007 from the 2001 archive data of the Parkes 64-m telescope. In 2013, Thornton et al. [6] discovered several other similar signals, thereafter FRBs began to attract people's interest. FRBs can be generally categorized into two classes: repeaters and non-repeaters. However, it is unclear whether there are essential differences between them. The first discovered repeating FRB is FRB 121102, which is extremely active and can repeat over a thousand times in $ \sim 60 $ hours [7]. Some other repeating FRBs might repeat only two or three times [8]. FRBs also have a high occurrence rate across the full sky. Until now, hundreds of FRBs have been discovered by various telescopes [1, 9], FRB 200428 confirmed that at least some FRBs may originate from magnetars, but there is still no consensus on the origin of most FRBs. The large dispersion measures (DMs) imply that the majority of FRBs have an extragalactic origin, with only one FRB confirmed to come from the Milky Way [10]. The localization of the host galaxy and the direct measurement of redshift confirm that they have an extragalactic origin [1113], thus making FRBs a useful tool to study the universe.

      Recently, the Hubble tension problem, i.e. the contradiction between the Hubble constant values measured from cosmic microwave background (CMB) and Type Ia supernovae (SNe Ia) calls people's attention [1417]. New probes for the universe are necessary to resolve the Hubble tension problem, and FRBs may have this potential. The high energy of FRBs makes it possible to be detected at high redshift, and the high DM also makes them useful cosmological probes. For instance, Macquart et al. [18] used FRBs to constrain the missing baryons in the universe. Gao et al. [19] used FRB/GRB systems to constrain dark energy. Zhao et al. [20] used simulated FRB data to break the parameter degeneracies inherent in the CMB data. Qiu et al. [21] combined simulated FRB data with current CMB data to constrain the holographic dark energy (HDE) model and the Ricci dark energy (RDE) model. Zhao et al. [22] used unlocalized FRB data to constrain the Hubble constant. Zhang et al. [23] investigated how the upcoming FRB observations with Square Kilometre Array (SKA) can help developing cosmology, and showed that $ 10^6 $ FRB data can tightly constrain dark-energy equation of state parameters and the baryon density $ \Omega_{\rm{b}}h $. Bhattacharya et al. [24] used FRBs to explore the reionization history of the universe. Lin & Sang [25] used FRBs to test the anisotropy of the universe. Wu et al. [26] used FRBs to constrain the Hubble constant. Li et al. [27] used the strongly lensed FRBs to constrain cosmological parameters, such as the Hubble constant and the curvature of the Universe. Pearson et al. [28] pointed out that the strongly lensed repeating FRBs can be used to search for gravitational waves.

      There are some key issues in using FRBs as cosmological probes. First, although hundreds of FRBs have been measured, only a few of them have a direct measurement of redshift. Fortunately, many more FRBs can be detected in the future by telescopes such as the Canadian Hydrogen Intensity Mapping Experiment (CHIME) and the Five-hundred-meter Aperture Spherical Telescope (FAST), and some of them are expected to have direct redshift measurement [29, 30]. Second, the DM contribution from the host galaxy ($ \rm DM_{host} $) is not fully understood, and it may be significantly different for every FRBs. It is difficult to precisely model the $ \rm DM_{host} $ term. Some papers assume that the $ \rm DM_{host} $ term accommodates the evolution of the star formation rate history, so that it evolves with redshift [31]. However, Lin et al. [32] found that there is no strong evidence for the redshift evolution of $ \rm DM_{host} $ within the present data. A more reasonable way is to consider the probability distribution of $ \rm DM_{host} $ and marginalize over it. Based on theoretical considerations as well as numerical simulations, Macquart et al. [18] and Zhang et al. [33] showed that $ \rm DM_{host} $ follows a log-normal distribution. Finally, due to the fluctuation of the matter density, the DM contribution from the intergalactic medium ($ \rm DM_{IGM} $) may significantly deviate from the mean value. Therefore, using the mean value to estimate $ \rm DM_{IGM} $ may strongly bias the results. One usual way is to treat $ \rm DM_{IGM} $ as Gaussian distribution centered on the mean value. The Gaussian distribution is useful owing to the Gaussianity of structure on large scales, but there may also be large skew resulting from a few large-scale structures. Considering this, Macquart et al. [18] proposed a quasi-Gaussian distribution of $ \rm DM_{IGM} $ based on hydrodynamic simulations and theoretical motivation. Different treatments of $ \rm DM_{IGM} $ may affect the constraint on cosmological parameters.

      In this paper, we use 18 well-localized FRBs to constrain the Hubble constant $ H_0 $. The probability distributions of $ \rm DM_{IGM} $ and $ \rm DM_{host} $ are taken into consideration, and the free parameters that govern the probability distributions are constrained simultaneously with the Hubble constant. The rest parts of this paper is arranged as follows: In Section 2, we introduce the Bayesian method that is used to constrain cosmological parameters. In Section 3 we give the observational data and show the constraining results. In Section 4, we perform the Monte Carlo simulations to check the validity of this method and to compare the impacts of different probability distributions of $ \rm DM_{IGM} $. Finally, discussion and conclusions are given in Section 5.

    II.   METHODS
    • When FRB signal travels from the host galaxy to the earth, it experiences a time delay due to the interaction of electromagnetic waves with free electrons. The time delay depends on the frequency of electromagnetic waves. Measuring the time delay between photons of different frequencies allows us to determine the number density of electrons integrated along the wave path, referred to as the dispersion measure (DM). The total dispersion measure of an extragalactic FRB can be generally separated into four components [18],

      $ {\rm{DM_{obs}} = DM_{MW} + DM_{halo} + DM_{IGM}} + \frac{\rm{DM_{host}}}{1+z}, $

      (1)

      where $ \rm DM_{MW} $ comes from the contribution of Galactic interstellar medium, which can be estimated from the Galactic electron density models such as the NE2001 model [34]. The $ \rm DM_{halo} $ term is the contribution from Galactic halo, which is not fully constrained yet. Prochaska et al. [35] provided an estimation that it is about $ \rm 50- 100\ pc\ cm^{-3} $. Here we follow Macquart et al. [18] and assume $ \rm DM_{halo}=50\ pc\ cm^{-3} $. The $ \rm DM_{host} $ term is the contribution from the host galaxy, and the factor $ 1+z $ arises from the cosmic expansion. The $ \rm DM_{IGM} $ term is the contribution from intergalactic medium, and it carries the information of the universe. Several other studies also mentioned the $ \rm DM_{source} $ term [3, 4, 23], which represents the DM contribution from the immediate environment of the source. $ \rm DM_{source} $ term is highly dependent on the source environment of each FRB, which is not fully understood yet. The value of $ \rm DM_{source} $ is usually smaller than the error of $ \rm DM_{host} $ and $ \rm DM_{IGM} $ [36]. Therefore, we ignored the $ \rm DM_{source} $ term here. The $ \rm DM_{MW} $ and $ \rm DM_{halo} $ terms can be subtracted from the total observed $ {\rm{DM}} $, leaving behind the extragalactic $ \rm DM $, which is defined by

      $ {\rm{DM_E\equiv DM_{obs}} - DM_{MW} - DM_{halo} = \rm DM_{IGM} } + \frac{\rm{DM_{host}}}{1+z}. $

      (2)

      In the flat $ \rm \Lambda CDM $ model, the average value of the $ \rm DM_{IGM} $ term can be written as [37, 38]

      $ \langle {\rm{DM_{IGM}}}(z)\rangle=\frac{3cH_0 {\rm{\Omega_b}} f_{\rm{IGM}} f_e}{8 \pi G m_p}\int_0^z \frac{1+z}{\sqrt{ {\rm{\Omega_m}}(1+z)^3+{\rm{\Omega_{\Lambda}}} }}\ dz, $

      (3)

      where $ H_0 $ is the Hubble constant, $ \rm \Omega_b $ is the cosmic baryon mass density, $ \Omega_m $ is the matter density, and $ \Omega_{\Lambda} $ is the vacuum energy density of the Universe. The c, G, and $ m_p $ are three constants, which represent the speed of light in vacuum, the Newtonian gravitational constant, and the mass of proton, respectively. $ f_e= Y_{\rm{H}} X_{e,{\rm{H}}}+\dfrac{1}{2}Y_{\rm{He}}X_{e,{\rm{He}}} $ denotes the extent of ionization progress of hydrogen and helium, where $ Y_{\rm{H}}=0.75 $ and $ Y_{\rm{He}}=0.25 $ are the hydrogen mass fraction and helium mass fraction respectively. $ X_{e,{\rm{H}}} $ and $ X_{e,{\rm{He}}} $ are the corresponding ionization fractions. Since both hydrogen and helium are expected to be fully ionized at $ z \leq 3 $ [39, 40], we take $ X_{e,{\rm{H}}}=X_{e,{\rm{He}}}=1 $. $ f_{\rm{IGM}} $ is the fraction of baryon mass in IGM.

      However, Eq. (3) only describes the mean value of $ \rm DM_{IGM} $. The actual value is associated with large-scale matter density fluctuations and may vary around the mean value. Usually, the Gaussian distribution is used to describe $ \rm DM_{IGM} $ [4143]. However, theoretical motivation and numerical simulations show that the probability distribution of $ \rm DM_{IGM} $ can be modeled using the quasi-Gaussian distribution [18, 44],

      $ p_{\rm{IGM}}(\Delta)=A\Delta^{-\beta}{\exp } \left[ -\frac{(\Delta^{-\alpha}-C_0)^2}{2\alpha^2\sigma_{\rm{IGM}}^2}\right],\quad \Delta>0, $

      (4)

      where $ \rm \Delta \equiv DM_{IGM}/\langle DM_{IGM} \rangle $, $ \rm \sigma_{IGM} $ is the effective standard deviation, α and β are related to the inner density profile of gas in haloes. Hydrodynamic simulations show that $ \alpha=\beta=3 $ provides the best fit to the model [18, 44], hence we fix these two parameters. We also followed Macquart et al. [18] to parameterize the effective standard deviation as $ {\rm{\sigma_{IGM}}} = Fz^{-1/2} $, where F is a free parameter and z is the redshift of FRB source. Here, A is a normalization constant and $ C_0 $ is chosen to make sure that the mean of this distribution is unity. Note that the distribution of $ \rm DM_{IGM} $ is a function of z, implying that A and $ C_0 $ vary with z.

      For the $ \rm DM_{host} $ term, it may range from tens to hundreds of $ \rm pc\ cm^{-3} $. For example, Xu et al. [45] estimated that the $ \rm DM_{host} $ of FRB20201124A can range from 10 to 310 $ \rm pc\ cm^{-3} $, while Niu et al. [46] estimated that the $ \rm DM_{host} $ of FRB20190520B can be up to 900 $ \rm pc\ cm^{-3} $. To account for the large variation of $ \rm DM_{host} $, it is often modeled with the log-normal distribution [18, 33],

      $\begin{aligned}[b]& p_{\rm{host}}({\rm{DM_{host}} }|\mu,\sigma_{\rm{host}})\\=\;&\frac{1}{\sqrt{2\pi}{\rm{DM_{host}}}\sigma_{\rm{host}}}{\exp} \left[-\frac{(\ln\rm DM_{host}-\mu)^2}{2\sigma_{\rm{host}}^2}\right],\end{aligned} $

      (5)

      where μ and $ \sigma_{\rm{host}} $ are the mean and standard deviation of $ \rm \ln{DM_{host}} $, respectively. This log-normal distribution allows for the appearance of large values of $ \rm DM_{host} $. Generally, both μ and $ \sigma_{\rm{host}} $ might be redshift-dependent. However, Zhang et al. [33] showed that for non-repeating bursts they do not vary significantly with redshift. Lin et al. [32] also proved that there is no strong evidence for the redshift evolution of $ \rm DM_{host} $ within the present data. Hence, we follow Macquart et al. [18] and treat them as two constants.

      For that $ \rm DM_{IGM} $ term and $ \rm DM_{host} $ term are challenging to separate, we introduce the probability distribution of the extragalactic $ \rm DM $ as [18]

      $\begin{aligned}\\[-8pt] p_{E}({\rm{DM_{E}}}|z) = \int_0^{(1+z){\rm{DM_E}}} p_{\rm{IGM}}({\rm{DM_E}}-\frac{\rm{DM_{host}}}{1+z}|H_0, F) \, p_{\rm{host}}({\rm{DM_{host}}}|\mu,\sigma_{\rm{host}}) \, d{\rm{DM_{host}}}. \end{aligned}$

      (6)

      If a large sample of well-localized FRBs are observed, the joint likelihood function can be written as

      $ \begin{array}{*{20}{l}} \mathcal{L}({\rm{FRBs}}|H_0, \mu,\sigma_{\rm{host}},F)=\prod\limits_{i=1}^N p_{E}({\rm{DM_{E,\it i}}}|z_i), \end{array} $

      (7)

      where N is the number of FRBs. According to Bayesian theorem, the posterior probability density function of the free parameters is given by

      $ \begin{aligned}[b]& P(H_0, \mu, \sigma_{\rm{host}},F|{\rm{FRBs}}) \\\propto\;& \mathcal{L}({\rm{FRBs}}|H_0, \mu, \sigma_{\rm{host}}, F) \, P_{0}(H_0, \mu, \sigma_{\rm{host}}, F), \end{aligned} $

      (8)

      where $ P_0 $ is the prior probability function of the parameters.

    III.   DATA AND RESULTS
    • Till now, there are over 30 published well-localized extragalactic FRBs that have identified host galaxies and well-measured redshifts 1. Among them, FRB20200120E, FRB20190614D, FRB20190520B and FRB 20220319D are excluded. FRB20200120E is so near to the Milky Way that the peculiar velocity dominates the Hubble flow, resulting in a negative spectroscopic redshift $ z=-0.001 $ [47, 48]. FRB20190614D only has a photometric redshift $ z\approx 0.6 $ [49], with no available spectroscopic redshift. The $ {\rm{DM_{host}}} $ of FRB20190520B is estimated to be as large as $ 900\ \rm pc\ cm^{-3} $ [46], which is much larger than the normal FRBs. FRB20220319D is also excluded due to its total DM being lower than $ \rm DM_{MW} $ [50]. The remaining 35 FRBs all have well-measured spectroscopic redshifts, and their main properties are listed in Table 1. These FRBs will be used to constrain the cosmological parameters.

      FRBs RA Dec $ {\rm DM_{obs}} $[$ ^{\circ} $] $ {\rm DM_{MW}} $[$ {\rm pc cm^{-3}} $] $ {\rm DM_E} $[$ {\rm pc cm^{-3}} $] $ z_{\rm sp} $[$ {\rm pc cm^{-3}} $] reference
      20121102A 82.99 33.15 557.00 157.60 349.40 0.1927 Chatterjee et al. [12]
      20171020A 22.15 −19.40 114.10 38.00 26.10 0.0087 Li et al. [51]
      20180301A 93.23 4.67 536.00 136.53 349.47 0.3305 Bhandari et al. [52]
      20180916B 29.50 65.72 348.80 168.73 130.07 0.0337 Marcote et al. [53]
      20180924B 326.11 −40.90 362.16 41.45 270.71 0.3214 Bannister et al. [54]
      20181030A 158.60 73.76 103.50 40.16 13.34 0.0039 Bhardwaj et al. [55]
      20181112A 327.35 −52.97 589.00 41.98 497.02 0.4755 Prochaska et al. [56]
      20190102C 322.42 −79.48 364.55 56.22 258.33 0.2913 Macquart et al. [18]
      20190523A 207.06 72.47 760.80 36.74 674.06 0.6600 Ravi et al. [57]
      20190608B 334.02 −7.90 340.05 37.81 252..24 0.1178 Macquart et al. [18]
      20190611B 320.74 −79.40 332.63 56.60 226.03 0.3778 Macquart et al. [18]
      20190711A 329.42 −80.36 592.60 55.37 487.23 0.5217 Macquart et al. [18]
      20190714A 183.98 −13.02 504.13 38.00 416.13 0.2365 Heintz et al. [58]
      20191001A 323.35 −54.75 507.90 44.22 413.68 0.2340 Heintz et al. [58]
      20191228A 344.43 −29.59 297.50 33.75 213.75 0.2432 Bhandari et al. [52]
      20200430A 229.71 12.38 380.25 27.35 302.90 0.1608 Bhandari et al. [52]
      20200906A 53.50 −14.08 577.80 36.19 491.61 0.3688 Bhandari et al. [52]
      20201124A 77.01 26.06 413.52 126.49 237.03 0.0979 Fong et al. [59]
      20210405I 255.34 −48.48 565.17 468.18 46.99 0.066 Driessen et al. [60]
      20210410D 326.09 −78.68 572.62 56.20 466.45 0.1415 Caleb et al. [61]
      20210603A 10.27 21.23 500.15 40.00 410.15 0.1772 Cassanelli et al. [62]
      20211127A 199.81 −18.84 234.83 41.75 143.08 0.0496 Khrykin et al. [63]
      20211212A 157.35 1.36 206.00 38.29 117.71 0.0713 Khrykin et al. [63]
      20220207C 310.20 72.88 262.38 74.99 137.39 0.0430 Law et al. [64]
      20220307B 350.87 72.19 499.27 119.82 329.45 0.2481 Law et al. [64]
      20220310F 134.72 73.49 462.24 45.07 367.17 0.4780 Law et al. [64]
      20220418A 219.10 70.10 623.25 36.49 536.76 0.6220 Law et al. [64]
      20220506D 318.04 72.83 396.97 82.88 264.09 0.3004 Law et al. [64]
      20220509G 282.67 70.24 269.53 55.36 164.17 0.0894 Law et al. [64]
      20220825A 311.98 72.58 651.24 77.26 523.98 0.2414 Law et al. [64]
      20220912A 347.27 48.71 220.7 120.44 50.26 0.0771 Zhang et al. [65]
      20220914A 282.06 73.34 631.28 54.38 526.9 0.1139 Law et al. [64]
      20220920A 240.26 70.92 314.99 39.62 225.37 0.1582 Law et al. [64]
      20221012A 280.80 70.52 441.08 53.69 337.39 0.2847 Law et al. [64]
      20221022A 48.63 86.87 116.84 60.12 6.72 0.0149 Mckinven et al. [66]

      Table 1.  The properties of the Host/FRB catalog. Column 1: FRB name; Columns 2 and 3: the right ascension and declination of FRB source on the sky; Column 4: the observed DM; Column 5: the DM of the Milky Way ISM calculated using the NE2001 model; Column 6: the extragalactic DM calculated by subtracting $ {\rm{DM_{\rm MW}}} $ and $ {\rm{DM_{\rm halo}}} $ from the observed $ {\rm{DM_{\rm obs}}} $, assuming $ {\rm{DM_{\rm halo}}}=50\; {\rm{pc\; cm^{-3}}} $ for the Milky Way halo; Column 7: the spectroscopic redshift; Column 8: the references.

      We employ Monte Carlo Markov Chain (MCMC) analysis to constrain $ H_0 $ and three free parameters ($ e^{\mu},\ \sigma_{\rm{host}}, F $) from the probability functions of $ \rm DM_{host} $ and $ \rm DM_{IGM} $. We use $ e^{\mu} $ instead of μ as a free parameter because $ e^{\mu} $ directly represents the median value of $ \rm DM_{host} $. For $ f_{\rm{IGM}} $, there is no precise constraint yet. It may slowly increase with redshift [67], and Li et al. [31] parameterized it as $ f_{\rm{IGM}}=f_{{\rm{IGM}},0}(1+\alpha z/(1+z)) $. However, there is no strong evidence to show that $ f_{\rm{IGM}} $ increases with redshift in the current FRB data [68]. To be conservative, we assume that $ f_{\rm{IGM}} $ follows a uniform distribution $ U(0.747,0.913) $ and marginalized over it [26]. Additionally, We adopt $ \Omega_b h^2 = 0.0224 $ and $ \Omega_m = 0.315 $ from Planck 2018 results [69]. The posterior probability density functions of the free parameters are calculated using the publicly available Python code emcee [70]. Based on previous results [18, 71], flat priors are applied to all four parameters: $ H_0 \in U(0,100)\; {\rm{km\; s^{-1}}\; Mpc^{-1}} $, $ e^{\mu} \in U(20,200) {\rm{\ pc\ cm^{-3}}} $, $ \sigma_{\rm{host}} \in U(0.2,2) $, and $ F \in U(0.01,0.5) $.

      The 2D marginalized posterior distributions and the $ 1-3\ \sigma $ confidence contours of the four parameters are plotted in the left panel of Figure 1. It is evident that only the parameters $ H_0 $ and $ \sigma_{\rm{host}} $ are tightly constrained, while the other two parameters, $ e^{\mu} $ and F, are not well-constrained. One possible reason is that the FRB sample is not large enough to simultaneously constrain a model with too many free parameters. A notable feature is that the best-fitting value of the Hubble constant $ H_0=49.34^{+4.39}_{-3.83}\; {\rm{km\; s^{-1}}\; Mpc^{-1}} $ is much lower than the Planck 2018 value. This discrepancy may be caused by the correlation between parameters. As can be seen from the contour plot, $ H_0 $ seems to be positively correlated with $ e^{\mu}_{\rm{host}} $, while is negatively correlated with F.

      Figure 1.  Constraints on the four free parameters ($ H_0,\ e^{\mu},\ \sigma_{\rm{host}} $, F) and three free parameters ($ H_0,\ e^{\mu},\ \sigma_{\rm{host}} $) using the 35 FRBs samples. The black-dashed lines from left to right in each subfigure represents the 16%, 50% and 84% quantiles of the distribution, respectively. The contours from the inner to outer represent $ 1\sigma $, $ 2\sigma $ and $ 3\sigma $ confidence regions, respectively. For simplicity, units are omitted in this figure.

      Several works suggest that the parameter F prefers a value around $ F=0.2 $ or even smaller at $ z\leq 1 $ [72, 73]. However, our four-parameter fit from the FRB sample indicates a preference for a larger value of F, although it can not be tightly constrained. This discrepancy might introduce bias to the other parameters. To address this, we try to fix $ F = 0.2 $ while keeping the other three parameters $ (H_0,\ e^{\mu},\ \sigma_{\rm{host}}) $ free. The corresponding 2D marginalized posterior distributions and the $ 1-3\ \sigma $ confidence contours of the three parameters are presented in the right panel of Figure 1. One can see that both $ H_0 $ and $ \sigma_{\rm{host}} $ are well-constrained, the parameter $ e^{\mu}_{\rm{host}} $ is also constrained but with a relatively small value. The best-fitting Hubble constant, $ H_0=60.99^{+4.57}_{-4.90}\; {\rm{km\; s^{-1}}\; Mpc^{-1}} $, is notably larger than that from the four-parametric fit. However, the median value of the Hubble constant is still relatively lower than the Planck 2018 result. One possible explanation is that the FRB sample is not large enough to provide a robust result. The results of the two parameters from the probability functions of $ \rm DM_{host} $ are given by $ e^{\mu}_{\rm{host}}=48.35^{+20.81}_{-15.00} {\rm{\ pc\ cm^{-3}}} $ and $ \sigma_{\rm{host}}=1.63^{+0.23}_{-0.20} $, both of which have relatively large uncertainties.

    IV.   MONTE CARLO SIMULATION
    • With the progress of observational technique, more and more FRBs are expected to be detected, and a fraction of them can be well-localized. Therefore, it is interesting to assess the constraining capability of a large sample of FRBs across a wide redshift range. To this end, we conduct Monte Carlo simulations to investigate the efficiency of our method.

      The intrinsic redshift distribution of FRBs is still unclear due to the small well-localized sample. Li et al. [31] assumed that FRBs have a constant comoving number density but with a Gaussian cutoff. Zhang et al. [38] argued that the redshift distribution of FRBs is expected to be related with cosmic star formation rate (SFR), or influenced by the compact star merger but with an additional time delay. In this paper, we adopt the SFR-related model, in which the probability density function takes the form [38]

      $ P(z) \propto \frac{4\pi D_c^2(z){\rm{SFR}}(z)}{(1+z)H(z)}, $

      (9)

      where $ D_c(z)= \int_0^z \dfrac{cdz}{H(z)} $ represents the comoving distance, with c the speed of light and $ H(z) $ the Hubble expansion rate, and the SFR takes the form [74]

      $ {\rm{SFR}}(z)=0.02\left[(1+z)^{a\eta}+\left(\frac{1+z}{B}\right)^{b\eta}+\left(\frac{1+z}{C}\right)^{c\eta}\right]^{1/\eta} , $

      (10)

      where $ a=3.4,\ b=-0.3,\ c=-3.5,\ B=5000,\ C=9 $ and $ \eta = -10 $.

      We simulate a set of mock FRB samples to constrain the parameters. The simulations are performed based on the flat $ \rm \Lambda CDM $ model with Planck 2018 results [69]. The fiducial parameters include $ F=0.2,\ f_{\rm{IGM}}=0.83, \ e^{\mu}=100{\rm\ pc\ cm^{-3}} $ and $ \sigma_{\rm{host}}=1.0 $. The simulation procedures are outlined as follows. First, a certain number of redshifts are randomly drawn from Eq. (9), with $ z_{\rm{max}} $ set to $ 3.0 $. Subsequently, the same number of Δ values are randomly drawn from Eq. (4). The average $ \rm \langle DM_{IGM}(z)\rangle $ is calculated using Eq. (3), and $ \rm DM_{IGM} $ for each redshift is obtained as $ \Delta \times \langle {\rm{DM_{IGM}}}(z)\rangle $. Next, the same number of $ \rm DM_{host} $ values are randomly drawn from Eq. (5), and $ \rm DM_{E} $ is calculated according to Eq. (2). Finally there is a sample of mock FRBs $ (z_i, {\rm{DM_{E,\it i}}}) $.

      We use the mock FRBs to constrain the four parameters ($ H_0 $, $ e^\mu $, $ \sigma_{\rm{host}} $, F) using the same method described above. The corresponding contour plots constrained from $ N=100 $ mock FRBs are shown in the left panel of Figure 2. One can see that except for the parameter $ H_0 $, the other parameters can not be tightly constrained. Additionally, the parameter $ H_0 $ fails to recover the fiducial value within 1σ uncertainty.

      Figure 2.  Constraints on the four free parameters ($ H_0 $, $ e^\mu $, $ \sigma_{\rm{host}} $, F) using 100 mock FRBs with quasi-Gaussian distribution (left panel) and Gaussian distribution (right panel) of $ \rm DM_{IGM} $. The blue lines represent the fiducial values. The black-dashed lines from left to right in each subfigure represents the 16%, 50% and 84% quantiles of the distribution, respectively. The contours from the inner to outer represent $ 1\sigma $, $ 2\sigma $ and $ 3\sigma $ confidence regions, respectively. For simplicity, units are omitted in this figure.

      For comparison, we also conducted MCMC analysis using a Gaussian distribution of $ \rm DM_{IGM} $. The method is similar to the quasi-Gaussian distribution case, with the primary difference being the replacement of the distribution function of $ \rm DM_{IGM} $ (i.e., Eq. (4)) with the Gaussian distribution $ \mathcal{G}(\langle {\rm{DM_{IGM}}} \rangle, \sigma_{\rm{IGM}}) $ in generating mock FRBs samples, as well as in the joint likelihood function. In the simulation, the fiducial value of $ \sigma_{\rm{IGM}} $ is set to $ 100{\rm{\ pc\ cm^{-3}}} $ [41, 42]. In the MCMC analysis, $ \sigma_{\rm{IGM}} $ is treated as a free parameter, replacing the parameter F of the quasi-Gaussian case. And the prior of $ \sigma_{\rm{IGM}} $ is $ U(0,200)\ {\rm{pc\ cm^{-3}}} $. We employ mock FRBs generated with a Gaussian distribution of $ \rm DM_{IGM} $ to constrain the four parameters ($ H_0 $, $ e^\mu $, $ \sigma_{\rm{host}} $, $ \sigma_{\rm{IGM}} $). The corresponding contour plots constrained from $ N=100 $ mock FRBs are shown in the right panel of Figure 2. We can see that with the Gaussian distribution of $ \rm DM_{IGM} $, the parameters $ H_0 $ and $ \sigma_{\rm{host}} $ can well recover their fiducial values. However, the other two parameters ($ e^{\mu} $, $ \sigma_{\rm{IGM}} $), especially $ \sigma_{\rm{IGM}} $, can not be tightly constrained, showing a strong bias with respect to the fiducial value.

      Since the best-fitting value of F is much larger than expected, we then fix $ F = 0.2 $ to see if the results can be improved or not. The corresponding contour plots constrained from $ N=100 $ mock FRBs with quasi-Gaussian distribution of $ \rm DM_{IGM} $ are shown in the left panel of Figure 3. With 100 mock FRBs, we obtain a well-constrained Hubble constant, providing a more precise result compared to the situation with 35 real FRBs. The other two parameters, $ e^\mu $ and $ \sigma_{\rm{host}} $, are also well-constrained but exhibit relatively large uncertainty. Regardless of the large uncertainty, all three parameters can well recover the fiducial values within $ 1\sigma $ uncertainty. However, with a thousand repetitions, each involving a sample of 100 mock FRBs, nearly half of the estimates for $ e^{\mu} $ and $ \sigma_{\rm{host}} $ could not be well constrained. Therefore, caution is advised when discussing the results of $ e^{\mu} $ and $ \sigma_{\rm{host}} $ in real-world situations.

      Figure 3.  Constraints on the three free parameters ($ H_0 $, $ e^\mu $, $ \sigma_{\rm{host}} $) using 100 mock FRBs with quasi-Gaussian distribution (left panel) and Gaussian distribution (right panel) of $ \rm DM_{IGM} $. The blue lines represent the fiducial values. The black-dashed lines from left to right in each subfigure represents the 16%, 50% and 84% quantiles of the distribution, respectively. The contours from the inner to outer represent $ 1\sigma $, $ 2\sigma $ and $ 3\sigma $ confidence regions, respectively. For simplicity, units are omitted in this figure.

      For comparison, the contour plots corresponding to a Gaussian distribution of $ \rm DM_{IGM} $ constrained from $ N=100 $ mock FRBs are also presented in the right panel of Figure 3. We can see that the quasi-Gaussian distribution of $ \rm DM_{IGM} $ may influence the results of $ e^{\mu} $ and $ \sigma_{\rm{host}} $. However, for the Hubble constant, the quasi-Gaussian distribution of $ \rm DM_{IGM} $ does not affect its reliability, it yields consistent results. We also do this comparison with $ N=500 $ mock FRBs, and the results are presented in Figure 4. For the Gaussian distribution of $ \rm DM_{IGM} $, all three parameters are well-constrained and can recover the fiducial values. In the case of the quasi-Gaussian distribution, the parameter $ e^{\mu} $ is not well-constrained, as mentioned before, due to the influence fo the $ \rm DM_{IGM} $ distribution. The Hubble constant $ H_0 $ can still be constrained and recover the fiducial value within $ 1\sigma $ uncertainty, albeit with a relatively small value. Furthermore, the results indicate that further enlarging the FRB sample size does not significantly improve the precision of constraint on $ H_0 $, primarily due to the uncertainty associated with $ f_{\rm{IGM}} $.

      Figure 4.  Constraints on the three free parameters ($ H_0 $, $ e^\mu $, $ \sigma_{\rm{host}} $) using 500 mock FRBs with quasi-Gaussian distribution (left panel) and Gaussian distribution (right panel) of $ \rm DM_{IGM} $. The blue lines represent the fiducial values. The black-dashed lines from left to right in each subfigure represents the 16%, 50% and 84% quantiles of the distribution, respectively. The contours from the inner to outer represent $ 1\sigma $, $ 2\sigma $ and $ 3\sigma $ confidence regions, respectively. For simplicity, units are omitted in this figure.

      To mitigate simulation fluctuations, we perform the simulation 100 times. Specifically, we randomly generate 100 FRB samples, with each sample containing 100 mock FRBs. The 100 mock samples are then used to constrain the three free parameters ($ H_0 $, $ e^\mu $, $ \sigma_{\rm{host}} $) using the method described above, resulting in 100 sets of best-fitting parameters. The distributions of the best-fitting parameter $ H_0 $ in the 100 simulations are shown in the left panel of Figure 5. The fiducial values are also shown as the red-solid lines. We can see that the Hubble constant can be tightly constrained in all cases. For comparison, we also present the results with a Gaussian distribution of $ \rm DM_{IGM} $ in the right panel of Figure 5. It is obvious that with the quasi-Gaussian distribution of $ \rm DM_{IGM} $, the best-fitting value of $ H_0 $ is systematically lower than the fiducial value, although it is still consistent with the fiducial value within $ 1\sigma $ uncertainty. On the contrary, if $ \rm DM_{IGM} $ is modelled with Gaussian distribution, the best-fitting value of $ H_0 $ well recovers the fiducial value. This implies that the choice of the distribution of $ \rm DM_{IGM} $ may cause bias on the estimation of Hubble constant.

      Figure 5.  (color online) The distribution of the best-fitting value for the Hubble constant is shown with quasi-Gaussian distribution (left panel) and Gaussian distribution (right panel) of $ \rm DM_{IGM} $ in 100 simulations, with $ N=100 $ FRBs in each simulation. The black dots and blue lines represent the median values of Hubble constant and their corresponding 1σ uncertainties, respectively. The red lines represent the fiducial values.

    V.   DISCUSSION AND CONCLUSIONS
    • In this paper, we employ Bayesian inference method to constrain the Hubble constant $ H_0 $, together with three FRB-related parameters ($ e^\mu $, $ \sigma_{\rm{host}} $, F) using 35 well-localized FRBs. We find that 35 FRBs couldn't tightly constrain four parameters simultaneously. Specifically, although the Hubble constant could be constrained, the best-fitting value $ H_0=49.34^{+4.39}_{-3.83}\; {\rm{km\; s^{-1}}\; Mpc^{-1}} $ is significantly lower than that measured from CMB. Additionally, the FRB-related parameter F is found to be much larger than expected. Subsequently, we addressed this discrepancy by fixing the parameter $ F=0.2 $ based on other observations. The three-parameter fit yielded $ H_0=60.99^{+4.57}_{-4.90}\ {\rm{km\ s^{-1}}\ Mpc^{-1}} $, a result that is still smaller than the results from CMB measurements. The relatively small median value of the Hubble constant may be attributed to the limited size of the FRB sample, or the correlation between $ H_0 $ and the FRB-related parameters.

      Considering the anticipated growth in the number of well-located FRBs in the future, we conducted Monte Carlo simulations to assess the efficiency of our method. It is found that for 100 mock FRBs, $ H_0 $ was effectively constrained, but the best-fitting $ H_0 $ is still lower than the fiducial value. This situation persists even if the number of FRBs increases to 500, and the precision of the results did not significantly improve due to the uncertainty of $ f_{\rm{IGM}} $. Further simulations reveal a systematical bias in the estimation of Hubble constant using FRBs, although the best-fitting value of $ H_0 $ is consistent with the fiducial value within $ 1\sigma $ uncertainty. This is mainly due to the quasi-Gaussian distribution of $ \rm DM_{IGM} $, which is correlated with the Hubble constant. The bias in $ \rm DM_{IGM} $ will cause a biased estimation on $ H_0 $.

      The quasi-Gaussian distribution of $ \rm DM_{IGM} $ accounts for the large skew resulting from a few large-scale structures in the universe. However, the non-Gaussianity may impact the fitting result of cosmological parameters, especially the Hubble constant. To prove our hypothesis, we simulated a set of FRB samples whose $ \rm DM_{IGM} $ values are drawn from Gaussian distribution. We then use the mock sample to constrain the Hubble constant, and compared the results with the quasi-Gaussian case in Figure 5. We find that in the Gaussian case, the best-fitting $ H_0 $ correctly recovers the fiducial value. However, in the quasi-Gaussian case, the estimation on $ H_0 $ is significantly biased, towards a systematically lower value than the fiducial value. Recognizing and addressing this bias is crucial for refining the fitting results of the Hubble constant.

      Recently, Wu et al. [26] have also used well-localized FRBs to constrain $ H_0 $ employing a quasi-Gaussian distribution of $ \rm DM_{IGM} $. They obtained results of $ H_0=68.81^{+4.99}_{-4.33}\; {\rm{km\; s^{-1}}\; Mpc^{-1}} $ if $ f_{\rm{IGM}}=0.83 $ is fixed, and $ H_0=69.31^{+6.21}_{-6.63}\; {\rm{km\; s^{-1}}\; Mpc^{-1}} $ if $ f_{\rm{IGM}} $ follows a uniform distribution $ U(0.747,0.913) $. The median values of both results are relatively higher than ours. The reasons for this difference are diverse. First, the FRB sample we used here are much larger than theirs. Second, for other cosmological parameters, such as $ \Omega_bh^2 $ and $ \Omega_m $, Wu et al. [26] treated them as free parameters, but with a uniform prior limited to a very small range. This is approximately equivalent to fix these two parameters, as is done in our work. Third, for the $ \rm DM_{halo} $ term, we set it as a constant, while Wu et al. [26] considered a Gaussian distribution to describe it. Finally, the main difference is the treatment of FRB-related parameters. For parameters (A, $ C_0 $, $ e^{\mu},\ \sigma_{\rm{host}} $, F) in the probability functions of $ \rm DM_{host} $ and $ \rm DM_{IGM} $, Wu et al. [26] chosen them based on the results of the state-of-the-art IllustrisTNG simulation. This approach may introduce a loop problem due to the fiducial parameter settings in the IllustrisTNG simulation. The method of Wu et al. [26] is equivalent to fix the FRB-related parameters in the MCMC analysis. While in our work, the FRB-related parameters are free. Since the FRB-related parameters are correlated with the Hubble constant, the estimation on $ H_0 $ in our work is biased and has large uncertainty.

Reference (74)

目录

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return