Skip to main content
Log in

Galaxy cluster number count data constraints on cosmological parameters

  • Regular Article - Theoretical Physics
  • Published:
The European Physical Journal C Aims and scope Submit manuscript

Abstract

We use data on massive galaxy clusters (M cluster>8×1014 h −1 M within a comoving radius of R cluster=1.5h −1 Mpc) in the redshift range 0.05≲z≲0.83 to place constraints, simultaneously, on the nonrelativistic matter density parameter Ω m , on the amplitude of mass fluctuations σ 8, on the index n of the power-law spectrum of the density perturbations, and on the Hubble constant H 0, as well as on the equation-of-state parameters (w 0,w a ) of a smooth dark energy component.

For the first time, we properly take into account the dependence on redshift and cosmology of the quantities related to cluster physics: the critical density contrast, the growth factor, the mass conversion factor, the virial overdensity, the virial radius and, most importantly, the cluster number count derived from the observational temperature data.

We show that, contrary to previous analyses, cluster data alone prefer low values of the amplitude of mass fluctuations, σ 8≤0.69 (1σ C.L.), and large amounts of nonrelativistic matter, Ω m ≥0.38 (1σ C.L.), in slight tension with the ΛCDM concordance cosmological model, though the results are compatible with ΛCDM at 2σ. In addition, we derive a σ 8 normalization relation, \(\sigma_{8} \varOmega_{m}^{1/3} = 0.49 \pm 0.06\) (2σ C.L.).

Combining cluster data with σ 8-independent baryon acoustic oscillation observations, cosmic microwave background data, Hubble constant measurements, Hubble parameter determination from passively evolving red galaxies, and magnitude–redshift data of type Ia supernovae, we find \(\varOmega_{m} = 0.28^{+0.03}_{-0.02}\) and \(\sigma_{8} = 0.73^{+0.03}_{-0.03}\), the former in agreement and the latter being slightly lower than the corresponding values in the concordance cosmological model. We also find \(H_{0} = 69.1^{+1.3}_{-1.5}~\mbox {km}/\mbox {s}/\mbox {Mpc}\), the fit to the data being almost independent on n in the adopted range [0.90,1.05].

Concerning the dark energy equation-of-state parameters, we show that the present data on massive clusters weakly constrain (w 0,w a ) around the values corresponding to a cosmological constant, i.e. (w 0,w a )=(−1,0). The global analysis gives \(w_{0} = -1.14^{+0.14}_{-0.16}\) and \(w_{a} = 0.85^{+0.42}_{-0.60}\) (1σ C.L. errors). Very similar results are found in the case of time-evolving dark energy with a constant equation-of-state parameter w=const (the XCDM parametrization). Finally, we show that the impact of bounds on (w 0,w a ) is to favor top-down phantom models of evolving dark energy.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. It has been known for some time that a spatially flat ΛCDM model is consistent with most observational constraints, see, e.g., Refs. [1113]. In the ΛCDM model the energy budget is dominated by far by a cosmological constant, a spatially homogeneous fluid with equation-of-state parameter w Λ =p Λ /ρ Λ =−1 (where p Λ and ρ Λ are the fluid pressure and energy density), with nonrelativistic CDM being the second largest contributor. Note that the “standard” CDM structure formation model—which the standard ΛCDM cosmological model assumes—might have some observational inconsistencies (see, e.g., [14, 15]).

  2. In dynamical dark energy models the dark energy density decreases in time and so remains comparable to the nonrelativistic matter (CDM and baryons) energy density for a longer time (than does a time-independent Λ). This partially alleviates the “coincidence” puzzle of the ΛCDM model. In addition, some dynamical dark energy scalar-field models have a nonlinear attractor solution that generates the current, tiny, dark energy density energy scale of order an meV from a significantly higher energy density scale (possibly of a more fundamental model) as a consequence of the very slow decrease in time of the dark energy density during the very long age of the Universe. These results are often viewed as providing significant theoretical motivation to consider dynamical dark energy models [16, 17].

  3. This is consistent with the simplest interpretation of the CMB anisotropy data, see, e.g. [33, 34].

  4. As anticipated, the normalization constant A can be related to the other cosmological parameters, Ω b , Ω c , h, and n: \(A = 2 \pi^{2} \sigma_{8}^{2} R_{8}^{n+3} / \int_{0}^{\infty}dy\,y^{n+2} T^{2}(y/R_{8}) W^{2}(y)\).

  5. Defining r Δ as the physical radius containing an overdensity of \(\varDelta_{v}'\) relative to the critical density, and M(r<r Δ ) as the mass contained in r Δ , the mass–temperature relation (32) assumes the standard \(T_{X}^{3/2}\) power-law form [43, 44], namely \(M(r < r_{\varDelta}) = \kappa_{\varDelta}^{3/2} [3/4\pi \varDelta'_{v} \rho_{\rm cr}^{(0)} ]^{1/2} (1+z)^{-3/2} T_{X}^{3/2}\).

  6. See Refs. [4749] for recent discussions of BAO data constraints on cosmological parameters.

  7. Since the upper limit of integration of the integral in Eq. (46) is infinity we must include, in the expression of the normalized Hubble parameter E(z), the contribution due to radiation. Accordingly, we add the quantity \(\rho_{r}/\rho_{\rm cr}^{(0)} = \varOmega_{r} (1+z)^{4}\) in the argument of the square root appearing in Eq. (5), where ρ r and Ω r are the radiation energy density and radiation density parameter, respectively. We take Ω r h 2=4.31×10−5 [51].

  8. This is reasonably consistent with both ‘low’ [55] and ‘high’ [56] recent estimates of the Hubble constant.

  9. See Refs. [6065] for Hubble parameter measurement constraints on cosmological parameters.

  10. This classification is not exhaustive since, as explained in [93], some models, such as those with oscillating equations of state, do not fall into it.

  11. There are a few disadvantages of the parametrization of Eq. (1): (i) This two parameter (w 0,w a ) parametrization has one more parameter than the simplest consistent and physically motivated scalar-field dark energy model [16, 17], thus making it more difficult to constrain model parameters with observational data; and, (ii) even so, the parametrization is not physically complete as additional information must be provided if one is interested in the evolution of energy density and other spatial inhomogeneities [101, 102].

  12. To be precise, B and β can (weakly) depend on the cosmology and the values used here are those found in Ref. [106] where a cosmology with cosmological constant was assumed.

  13. Here, we are considering the case where the dark energy does not cluster and does not virialize, so that the only terms in the potential energy which are relevant for energy conservation and virialization are those in Eq. (C.8) [107].

  14. It is straightforward to show that the only case where U mDE(r,z) is z-independent, and the system is conservative, is that of the cosmological constant, namely w(z)=−1 for all times.

References

  1. S.W. Allen, A.E. Evrard, A.B. Mantz, arXiv:1103.4829 [astro-ph.CO]

  2. P.J.E. Peebles, Astrophys. J. 284, 439 (1984)

    Article  ADS  Google Scholar 

  3. S. Weinberg, Cosmology (Oxford University Press, New York, 2008)

    MATH  Google Scholar 

  4. G.M. Voit, Rev. Mod. Phys. 77, 207 (2005). arXiv:astro-ph/0410173

    Article  ADS  Google Scholar 

  5. P. Brax, arXiv:0912.3610 [astro-ph.CO]

  6. J.-P. Uzan, arXiv:0912.5452 [gr-qc]

  7. A. De Felice, S. Tsujikawa, Living Rev. Relativ. 13, 3 (2010). arXiv:1002.4928 [gr-qc]

    ADS  Google Scholar 

  8. E.V. Linder, arXiv:1004.4646 [astro-ph.CO]

  9. A. Blanchard, Astron. Astrophys. Rev. 18, 595 (2010). arXiv:1005.3765 [astro-ph.CO]

    Article  ADS  MathSciNet  Google Scholar 

  10. D. Sapone, Int. J. Mod. Phys. A 25, 5253 (2010). arXiv:1006.5694 [astro-ph.CO]

    Article  ADS  MATH  Google Scholar 

  11. H.K. Jassal, J.S. Bagla, T. Padmanabhan, Mon. Not. R. Astron. Soc. 405, 2639 (2010). arXiv:astro-ph/0601389

    ADS  Google Scholar 

  12. K.M. Wilson, G. Chen, B. Ratra, Mod. Phys. Lett. A 21, 2197 (2006). arXiv:astro-ph/0602321

    Article  ADS  Google Scholar 

  13. T.M. Davis et al., Astrophys. J. 666, 716 (2007). arXiv:astro-ph/0701510

    Article  ADS  Google Scholar 

  14. P.J.E. Peebles, B. Ratra, Rev. Mod. Phys. 75, 559 (2003). arXiv:astro-ph/0207347

    Article  ADS  MATH  MathSciNet  Google Scholar 

  15. L. Perivolaropoulos, J. Phys. Conf. Ser. 222, 012024 (2010). arXiv:1002.3030 [astro-ph.CO]

    Article  ADS  Google Scholar 

  16. P.J.E. Peebles, B. Ratra, Astrophys. J. Lett. 325, L17 (1988)

    Article  ADS  Google Scholar 

  17. B. Ratra, P.J.E. Peebles, Phys. Rev. D 37, 3406 (1988)

    Article  ADS  Google Scholar 

  18. A. Mantz, S.W. Allen, H. Ebeling, D. Rapetti, Mon. Not. R. Astron. Soc. 387, 1179 (2008). arXiv:0709.4294 [astro-ph]

    Article  ADS  Google Scholar 

  19. A. Mantz, S.W. Allen, D. Rapetti, H. Ebeling, Mon. Not. R. Astron. Soc. 406, 1759 (2010). arXiv:0909.3098 [astro-ph.CO]

    ADS  Google Scholar 

  20. A. Vikhlinin et al., Astrophys. J. 692, 1060 (2009). arXiv:0812.2720 [astro-ph]

    Article  ADS  Google Scholar 

  21. N.A. Arhipova, T. Kahniashvili, V.N. Lukash, Astron. Astrophys. 386, 775 (2002). arXiv:astro-ph/0110426

    Article  ADS  MATH  Google Scholar 

  22. T. Kahniashvili, E. von Toerne, N.A. Arhipova, B. Ratra, Phys. Rev. D 71, 125009 (2005). arXiv:astro-ph/0503328

    Article  ADS  Google Scholar 

  23. E. Rozo et al., Astrophys. J. 708, 645 (2010). arXiv:0902.3702 [astro-ph.CO]

    Article  ADS  Google Scholar 

  24. S. Basilakos, M. Plionis, J.A.S. Lima, Phys. Rev. D 82, 083517 (2010). arXiv:1006.3418 [astro-ph.CO]

    Article  ADS  Google Scholar 

  25. M. Manera, D.F. Mota, Mon. Not. R. Astron. Soc. 371, 1373 (2006). astro-ph/0504519

    Article  ADS  Google Scholar 

  26. S. Basilakos, M. Plionis, J. Sola, Phys. Rev. D 80, 083511 (2009). arXiv:0907.4555 [astro-ph.CO]

    Article  ADS  Google Scholar 

  27. J. Grande, J. Sola, S. Basilakos, M. Plionis, J. Cosmol. Astropart. Phys. 1108, 007 (2011). arXiv:1103.4632 [astro-ph.CO]

    Article  ADS  Google Scholar 

  28. N.A. Bahcall, X.-h. Fan, Astrophys. J. 504, 1 (1998). arXiv:astro-ph/9803277

    Article  ADS  Google Scholar 

  29. N.A. Bahcall, P. Bode, Astrophys. J. 588, L1 (2003). arXiv:astro-ph/0212363

    Article  ADS  Google Scholar 

  30. M. Chevallier, D. Polarski, Int. J. Mod. Phys. D 10, 213 (2001). arXiv:gr-qc/0009008

    Article  ADS  Google Scholar 

  31. E.V. Linder, Phys. Rev. Lett. 90, 091301 (2003). arXiv:astro-ph/0208512

    Article  ADS  Google Scholar 

  32. L.R. Abramo, R.C. Batista, R. Rosenfeld, J. Cosmol. Astropart. Phys. 0907, 040 (2009). arXiv:0902.3226 [astro-ph.CO]

    Article  ADS  Google Scholar 

  33. S. Podariu, T. Souradeep, J.R. Gott, B. Ratra, M.S. Vogeley, Astrophys. J. 559, 9 (2001). arXiv:astro-ph/0102264

    Article  ADS  Google Scholar 

  34. E. Komatsu et al. (WMAP Collaboration), Astrophys. J. Suppl. Ser. 192, 18 (2011). arXiv:1001.4538 [astro-ph.CO]

    Article  ADS  Google Scholar 

  35. W.H. Press, P. Schechter, Astrophys. J. 187, 425 (1974)

    Article  ADS  Google Scholar 

  36. R.K. Sheth, G. Tormen, Mon. Not. R. Astron. Soc. 308, 119 (1999). arXiv:astro-ph/9901122

    Article  ADS  Google Scholar 

  37. D.J. Eisenstein, W. Hu, Astrophys. J. 496, 605 (1998). arXiv:astro-ph/9709112

    Article  ADS  Google Scholar 

  38. Y. Ikebe, T.H. Reiprich, H. Boehringer, Y. Tanaka, T. Kitayama, Astron. Astrophys. 383, 773 (2002). arXiv:astro-ph/0112315

    Article  ADS  Google Scholar 

  39. J.P. Henry, Astrophys. J. 534, 565 (2000). arXiv:astro-ph/0002365

    Article  ADS  Google Scholar 

  40. M. Schmidt, Astrophys. J. 151, 393 (1968)

    Article  ADS  Google Scholar 

  41. Y. Avni, J.N. Bahcall, Astrophys. J. 235, 694 (1980)

    Article  ADS  Google Scholar 

  42. M. Donahue, G.M. Voit, I.M. Gioia, G. Luppino, J.P. Hughes, J.T. Stocke, arXiv:astro-ph/9707010

  43. J. Hjorth, J. Oukbir, E. van Kampen, New Astron. Rev. 42, 145 (1998)

    Article  ADS  Google Scholar 

  44. J. Hjorth, J. Oukbir, E. van Kampen, Mon. Not. R. Astron. Soc. 298, L1 (1998). arXiv:astro-ph/9802293

    Article  ADS  Google Scholar 

  45. P.E.J. Nulsen, S.L. Powell, A. Vikhlinin, Astrophys. J. 722, 55 (2010). arXiv:1008.2393 [astro-ph.CO]

    Article  ADS  Google Scholar 

  46. G.L. Fogli, E. Lisi, A. Marrone, D. Montanino, A. Palazzo, Phys. Rev. D 66, 053010 (2002). arXiv:hep-ph/0206162

    Article  ADS  Google Scholar 

  47. E. Gaztañaga, A. Cabré, L. Hui, Mon. Not. R. Astron. Soc. 399, 1663 (2009). arXiv:0807.3551 [astro-ph]

    Article  ADS  Google Scholar 

  48. L. Samushia, B. Ratra, Astrophys. J. 701, 1373 (2009). arXiv:0810.2104 [astro-ph]

    Article  ADS  Google Scholar 

  49. Y. Wang, Mod. Phys. Lett. A 25, 3093 (2010), and references therein, arXiv:0904.2218 [astro-ph.CO]

    Article  ADS  MATH  Google Scholar 

  50. W.J. Percival et al. (SDSS Collaboration), Mon. Not. R. Astron. Soc. 401, 2148 (2010). arXiv:0907.1660 [astro-ph.CO]

    Article  ADS  Google Scholar 

  51. E.W. Kolb, M.S. Turner, The Early Universe (Addison-Wesley, Redwood City, 1990)

    Google Scholar 

  52. P.S. Corasaniti, A. Melchiorri, Phys. Rev. D 77, 103507 (2008). arXiv:0711.4119 [astro-ph]

    Article  ADS  Google Scholar 

  53. G. Chen, B. Ratra, Publ. Astron. Soc. Pac. 123, 1127 (2011). arXiv:1105.5206 [astro-ph.CO]

    Article  ADS  Google Scholar 

  54. G. Chen, J.R. Gott, B. Ratra, Publ. Astron. Soc. Pac. 115, 1269 (2003). arXiv:astro-ph/0308099

    Article  ADS  Google Scholar 

  55. G.A. Tamman, A. Sandage, B. Reindl, Astron. Astrophys. Rev. 15, 289 (2008). arXiv:0806.3018 [astro-ph]

    Article  ADS  Google Scholar 

  56. W.L. Freedman, B.F. Madore, Annu. Rev. Astron. Astrophys. 48, 673 (2010). arXiv:1004.1856 [astro-ph.CO]

    Article  ADS  Google Scholar 

  57. R. Jimenez, A. Loeb, Astrophys. J. 573, 37 (2002). arXiv:astro-ph/0106145

    Article  ADS  Google Scholar 

  58. D. Stern, R. Jimenez, L. Verde, M. Kamionkowski, S.A. Stanford, J. Cosmol. Astropart. Phys. 1002, 008 (2010). arXiv:0907.3149 [astro-ph.CO]

    Article  ADS  Google Scholar 

  59. E. Gaztañaga, A. Cabré, L. Hui, Mon. Not. R. Astron. Soc. 399, 1663 (2009). arXiv:0807.3551 [astro-ph]

    Article  ADS  Google Scholar 

  60. L. Samushia, B. Ratra, Astrophys. J. Lett. 650, L5 (2006). arXiv:astro-ph/0607301

    Article  ADS  Google Scholar 

  61. H. Zhang, Z.-H. Zhu, J. Cosmol. Astropart. Phys. 0803, 007 (2008). arXiv:astro-ph/0703245

    Article  ADS  Google Scholar 

  62. A.A. Sen, R.J. Scherrer, Phys. Lett. B 659, 457 (2008). arXiv:astro-ph/0703416

    Article  ADS  Google Scholar 

  63. L. Samushia, G. Chen, B. Ratra, arXiv:0706.1963 [astro-ph]

  64. N. Pan, Y. Gong, Y. Chen, Z.-H. Zhu, Class. Quantum Gravity 27, 155015 (2010). arXiv:1005.4249 [astro-ph.CO]

    Article  ADS  MathSciNet  Google Scholar 

  65. Y. Chen, B. Ratra, Phys. Lett. B 703, 406 (2011), and references therein. arXiv:1106.4294 [astro-ph.CO]

    Article  ADS  Google Scholar 

  66. R. Amanullah et al., Astrophys. J. 716, 712 (2010). arXiv:1004.1711 [astro-ph.CO]

    Article  ADS  Google Scholar 

  67. J. Guy et al., Astron. Astrophys. 466, 11 (2007). arXiv:astro-ph/0701828

    Article  ADS  Google Scholar 

  68. A. Lewis, S. Bridle, Phys. Rev. D 66, 103511 (2002). arXiv:astro-ph/0205436

    Article  ADS  Google Scholar 

  69. L. Campanelli, P. Cea, G.L. Fogli, A. Marrone, Phys. Rev. D 83, 103503 (2011). arXiv:1012.5596 [astro-ph.CO]

    Article  ADS  Google Scholar 

  70. K.-H. Chae, G. Chen, B. Ratra, D.-W. Lee, Astrophys. J. Lett. 607, L71 (2004). arXiv:astro-ph/0403256

    Article  ADS  Google Scholar 

  71. S. Lee, K.-W. Ng, Phys. Rev. D 76, 043518 (2007). arXiv:0707.1730 [astro-ph]

    Article  ADS  Google Scholar 

  72. M. Yashar, B. Bozek, A. Abrahamse, A. Albrecht, M. Barnard, Phys. Rev. D 79, 103004 (2009). arXiv:0811.2253 [astro-ph]

    Article  ADS  Google Scholar 

  73. M. Biesiada, A. Piorkowska, B. Malec, Mon. Not. R. Astron. Soc. 406, 1055 (2010), and references therein. arXiv:1105.0946 [astro-ph.CO]

    ADS  Google Scholar 

  74. S. Eidelman et al. (Particle Data Group), Phys. Lett. B 592, 1 (2004)

    Article  ADS  Google Scholar 

  75. P. Schuecker, H. Bohringer, C.A. Collins, L. Guzzo, Astron. Astrophys. 398, 867 (2003). arXiv:astro-ph/0208251

    Article  ADS  Google Scholar 

  76. A.G. Riess et al., Astrophys. J. 730, 119 (2011) [Erratum-ibid. 732, 129 (2011)]. arXiv:1103.2976 [astro-ph.CO]

    Article  ADS  Google Scholar 

  77. L.I. Gurvits, K.I. Kellermann, S. Frey, Astron. Astrophys. 342, 378 (1999). arXiv:astro-ph/9812018

    ADS  Google Scholar 

  78. E.J. Guerra, R.A. Daly, L. Wan, Astrophys. J. 544, 659 (2000). arXiv:astro-ph/0006454

    Article  ADS  Google Scholar 

  79. G. Chen, B. Ratra, Astrophys. J. 582, 586 (2003). arXiv:astro-ph/0207051

    Article  ADS  Google Scholar 

  80. M. Bonamente, M.K. Joy, S.J. La Roque, J.E. Carlstrom, E.D. Reese, K.S. Dawson, Astrophys. J. 647, 25 (2006). arXiv:astro-ph/0512349

    Article  ADS  Google Scholar 

  81. Y. Chen, B. Ratra, arXiv:1105.5660 [astro-ph.CO], and references therein

  82. S. Capozziello, V.F. Cardone, M. Funaro, S. Andreon, Phys. Rev. D 70, 123501 (2004). arXiv:astro-ph/0410268

    Article  ADS  Google Scholar 

  83. N. Pires, Z.-H. Zhu, J.S. Alcaniz, Phys. Rev. D 73, 123530 (2006). arXiv:astro-ph/0606689

    Article  ADS  Google Scholar 

  84. L. Samushia, A. Dev, D. Jain, B. Ratra, Phys. Lett. B 693, 509 (2010). arXiv:0906.2734 [astro-ph.CO]

    Article  ADS  Google Scholar 

  85. M.A. Dantas, J.S. Alcaniz, D. Mania, B. Ratra, Phys. Lett. B 699, 239 (2011), and references therein. arXiv:1010.0995 [astro-ph.CO]

    Article  ADS  Google Scholar 

  86. S.W. Allen, D.A. Rapetti, R.W. Schmidt, H. Ebeling, G. Morris, A.C. Fabian, Mon. Not. R. Astron. Soc. 383, 879 (2008). arXiv:0706.0033 [astro-ph]

    Article  ADS  Google Scholar 

  87. L. Samushia, B. Ratra, Astrophys. J. Lett. 680, L1 (2008). arXiv:0803.3775 [astro-ph]

    Article  ADS  Google Scholar 

  88. S. Ettori et al., Astron. Astrophys. 501, 61 (2009), and references therein. arXiv:0904.2740 [astro-ph.CO]

    Article  ADS  Google Scholar 

  89. B.E. Schaefer, Astrophys. J. 660, 16 (2007). arXiv:astro-ph/0612285

    Article  ADS  Google Scholar 

  90. N. Liang, S.N. Zhang, AIP Conf. Proc. 1065, 367 (2008). arXiv:0808.2655 [astro-ph]

    Article  ADS  Google Scholar 

  91. Y. Wang, Phys. Rev. D 78, 123532 (2008). arXiv:0809.0657 [astro-ph]

    Article  ADS  Google Scholar 

  92. L. Samushia, B. Ratra, Astrophys. J. 714, 1347 (2010), and references therein. arXiv:0905.3836 [astro-ph.CO]

    Article  ADS  Google Scholar 

  93. V. Barger, E. Guarnaccia, D. Marfatia, Phys. Lett. B 635, 61 (2006). arXiv:hep-ph/0512320

    Article  ADS  Google Scholar 

  94. R.R. Caldwell, E.V. Linder, Phys. Rev. Lett. 95, 141301 (2005). arXiv:astro-ph/0505494

    Article  ADS  Google Scholar 

  95. G. Gupta, S. Majumdar, A.A. Sen, Mon. Not. R. Astron. Soc. 420, 1309 (2012). arXiv:1109.4112 [astro-ph.CO]

    Article  ADS  Google Scholar 

  96. R.J. Scherrer, Phys. Rev. D 73, 043502 (2006). arXiv:astro-ph/0509890

    Article  ADS  MathSciNet  Google Scholar 

  97. C. Armendariz-Picon, V.F. Mukhanov, P.J. Steinhardt, Phys. Rev. Lett. 85, 4438 (2000). arXiv:astro-ph/0004134

    Article  ADS  Google Scholar 

  98. T. Chiba, Phys. Rev. D 73, 063501 (2006) [Erratum-ibid. D 80, 129901 (2009)]. arXiv:astro-ph/0510598

    Article  ADS  Google Scholar 

  99. A.Y. Kamenshchik, U. Moschella, V. Pasquier, Phys. Lett. B 511, 265 (2001). arXiv:gr-qc/0103004

    Article  ADS  MATH  Google Scholar 

  100. M.C. Bento, O. Bertolami, A.A. Sen, Phys. Rev. D 66, 043507 (2002). arXiv:gr-qc/0202064

    Article  ADS  Google Scholar 

  101. B. Ratra, Phys. Rev. D 43, 3802 (1991)

    Article  ADS  Google Scholar 

  102. S. Podariu, B. Ratra, Astrophys. J. 532, 109 (2000). arXiv:astro-ph/9910527

    Article  ADS  Google Scholar 

  103. F. Pace, J.C. Waizmann, M. Bartelmann, arXiv:1005.0233 [astro-ph.CO]

  104. D.F. Mota, J. Cosmol. Astropart. Phys. 0809, 006 (2008). arXiv:0812.4493 [astro-ph]

    Article  ADS  Google Scholar 

  105. J.F. Navarro, C.S. Frenk, S.D.M. White, Mon. Not. R. Astron. Soc. 275, 720 (1995). arXiv:astro-ph/9408069

    ADS  Google Scholar 

  106. J.S. Bullock et al., Mon. Not. R. Astron. Soc. 321, 559 (2001). arXiv:astro-ph/9908159

    Article  ADS  Google Scholar 

  107. P. Wang, Astrophys. J. 640, 18 (2006). arXiv:astro-ph/0507195

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We would like to thank A. Natarajan for useful comments. T.K. acknowledges partial support from Georgian National Science Foundation grant ST08/4-422, Swiss National Science Foundation SCOPES grant 128040, NASA grant NNXlOAC85G, NSF grant AST-1109180, the ICTP associate membership program, and the Università Degli Studi Di Bari for hospitality. B.R. acknowledges support from DOE grant DEFG03-99EP41093 and NSF grant AST-1109275.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to L. Campanelli.

Appendices

Appendix A: Critical density contrast and growth factor

Critical density contrast

The critical density contrast δ c depends on the redshift and on the cosmology and can be evaluated, using the approach of Ref. [103] (see also Ref. [104]), as follows. Consider the full nonlinear equation describing the evolution of the density contrast:

$$ \delta'' + \biggl( \frac{3}{a} + \frac{E'}{E} \biggr) \delta' - \frac{4}{3} \frac{\delta'^2}{1+\delta} - \frac{3}{2} \frac{\varOmega_m}{a^5 E^2} \delta(1+\delta) = 0 , $$
(A.1)

where a prime indicates differentiation with respect to the scale factor a and

$$ \delta = \frac{\delta \rho_m}{\rho_m} = \frac{\rho_{\mathrm{cluster}} - \rho_m}{\rho_m} $$
(A.2)

is the density contrast with ρ cluster being the cluster matter density. Since the above equation describes the nonlinear growth of the density contrast, its value at some chosen collapse time t collapse diverges. The critical density contrast δ c at the time t collapse is, by definition, the value of the density contrast at the time t collapse obtained by solving the linearized version of Eq. (A.1), namely

$$ \delta'' + \biggl( \frac{3}{a} + \frac{E'}{E} \biggr) \delta' - \frac{3}{2} \frac{\varOmega_m}{a^5 E^2} \delta = 0 , $$
(A.3)

with boundary conditions for δ such that the same boundary conditions applied to the nonlinear equation (A.1) make δ divergent at t collapse. Following Ref. [103], we take for the initial derivative of δ, δ′(a i ), the value δ′(a i )=5×10−5 where a i ≡5×10−5, while the initial value of the density contrast, δ(a i ), is found by searching for the value of δ(a i ) such that δ diverges at the time t collapse. We assume (as in Ref. [103]) that the divergency is achieved, numerically, when δ exceeds the value 107.

In Fig. 1, we plot the critical density contrast, δ c , as a function of the redshift for different values of (w 0,w a ). For large redshifts—where the effects of dark energy become negligible compared to those of nonrelativistic matter—the Universe effectively approaches the Einstein–de Sitter model where the critical density contrast is independent of the redshift and is δ c =(3/20)(12π)2/3≃1.686 [3]. In Fig. 7, we show the δ c -isocontours in the (w 0,w a )-plane for different values of the redshift and for Ω m =0.27.

Fig. 7
figure 7

δ c -isocontours in the (w 0,w a )-plane for different values of the redshift and for Ω m =0.27

Growth factor

The z-dependent part of the matter power spectrum—the growth factor D(z)—is

$$ D(z) = \frac{\delta(z)}{\delta(0)} , $$
(A.4)

and satisfies the linearized equation (A.3). The boundary conditions we impose are D(a=1)=1 and D(a=a i )=a i , where, as before, a i ≡5×10−5. In Fig. 1 we plot the growth factor, D(z), as a function of the redshift for different values of (w 0,w a ).

Appendix B: Mass conversion

The virial mass in the PS or ST parametrization needs to be expressed as a function of the observed mass M′ within a reference comoving radius of \(R_{0}' = 1.5 h^{-1}~\mbox {Mpc}\). As described in Sect. 2.2, what we really need is the fiducial virial mass M 0 as a function of the fiducial mass within \(R_{0}'\) adopted in the observation, namely \(M_{0}' = 8 \times 10^{14} h^{-1} M_{\odot}\). We use the following procedure to accomplish this. We first determine the physical virial radius r v,0, within which the virial mass M 0 is contained, through the relation

$$ M_0(z) = \frac{4\pi}{3} r_{v,0}^3(z) \rho_m(z) \varDelta_v(z) = \frac{4\pi}{3} R_{v,0}^3 \rho_m^{(0)} \varDelta_v(z) , $$
(B.1)

where R v,0=(1+z)r v,0, and Δ v is the virial overdensity and is discussed in Appendix C. We then scale the virial mass M 0 to the 1.5h −1 Mpc comoving radius assuming a Navarro–Frenk–White profile for the virialized halo mass density [105]:

$$ \rho_{\mathrm{cluster}}(r) = \frac{4\rho_{\mathrm{cluster}}(r_s)}{(r/r_s)(1+r/r_s)^2} , $$
(B.2)

where r is the physical radial distance and r s is a physical scale radius.

Technically the procedure is as follows. From Eq. (B.2) we can obtain the mass M(<r) contained in the physical radius r:

(B.3)

where

$$ f(x) = x^3 \biggl[ \ln \biggl( \frac{1+x}{x} \biggr) - \frac{1}{1+x} \biggr] . $$
(B.4)

Then applying Eq. (B.3) to the mass M 0 within the physical virial radius r v,0, and to the mass \(M_{0}'\) within the physical radius \(r'_{0} = R_{0}'/(1+z)\) corresponding to the comoving radius \(R'_{0}\), and taking the ratio of these two equations, we get

(B.5)

Here

$$ c_{v,0} = \frac{r_{v,0}}{r_s} $$
(B.6)

is the “concentration parameter”

$$ c_v = \frac{r_v}{r_s} \ $$
(B.7)

—the ratio between the physical virial radius r v and the physical scale radius r s —evaluated at the physical virial radius r v,0.

Both R v,0 and c v,0 in Eq. (B.5) are functions of M 0. From Eq. (B.1) we can express R v,0 as a function of M 0. For c v,0, we proceed as follows. First we consider the expression for the concentration parameter found by Bullock et al. [106] in their N-body simulation:

$$ c_v(M_v) = \frac{B}{1+z} \biggl[ \frac{M_v(0)}{M_*(0)} \biggr]^{-\beta} . $$
(B.8)

Here B≃9, β≃0.13,Footnote 12 M v is the mass within a physical virial radius r v ,

$$ M_v(z) = \frac{4\pi}{3} r_v^3(z) \rho_m(z) \varDelta_v(z) , $$
(B.9)

and M (z) is a fiducial mass defined by

$$ \sigma\bigl(R_*(z),z\bigr) = \delta_c(z) , $$
(B.10)

where the comoving radius R (z) is defined by

$$ M_*(z) = \frac{4\pi}{3} R_*^3(z) \rho_m^{(0)} . $$
(B.11)

Then, evaluating Eq. (B.8) for M v =M 0 [which corresponds to evaluating Eq. (B.7) for r v =r v,0], we have

$$ c_{v,0} = c_v(M_0) = \frac{c_{v,0}'}{\gamma} \biggl( \frac{M_0}{M_0'} \biggr)^{-\beta} , $$
(B.12)

where

(B.13)

is, formally, the concentration parameter (B.8) evaluated at \(M_{0}'\) [see Eq. (B.8)], and

(B.14)

Taking into account Eqs. (22), (24), and (B.10), we find that the quantity M (0) does not depend on M 0 and is defined by

$$ \varSigma\bigl(R_*(0)/R_0\bigr) = \frac{\delta_c(0)}{\sigma_8} . $$
(B.15)

Finally, inserting Eqs. (B.1) and (B.12) in Eq. (B.5), we obtain the equation that gives M 0 as a function of \(M_{0}'\) [namely the equation defining the function g in Eq. (31)]:

$$ f \biggl( R_{1.5} \frac{\gamma}{c_{v,0}'} \biggl( \frac{M_0}{M_0'} \biggr)^{\beta + 1/3} \biggr) = R_{1.5}^3 f \biggl( \frac{\gamma}{c_{v,0}'} \biggl( \frac{M_0}{M_0'} \biggr)^{\beta} \biggr) , $$
(B.16)

where

$$ R_{1.5}(z) = \frac{ [ 3M_0'/4\pi \rho_m^{(0)} \varDelta_v(z) ]^{1/3}}{1.5 h^{-1}~\mbox {Mpc}} $$
(B.17)

is the comoving virial radius corresponding to the mass \(M_{0}'\), normalized to 1.5h −1 Mpc.

In a flat universe dominated by nonrelativistic matter (the Einstein–de Sitter model), the comoving virial radius corresponding to the mass \(M_{0}' = 8 \times 10^{14} h^{-1} M_{\odot}\) is about 1.57h −1 Mpc, which gives R 1.5≃1. This in turn means, using Eq. (B.16), that \(M_{0}/M_{0}' \simeq 1\), namely the virial mass contained in a sphere of comoving virial radius of 1.5h −1 Mpc is approximatively \(M_{0}'\). Due to the presence of the third root in Eq. (B.17), the function R 1.5(z) is fairly insensitive to changes in the adopted cosmology [through \(\rho_{m}^{(0)}\) and Δ v (z)]. Therefore, also in the case of a generic cosmology with evolving dark energy, we expect values of M 0 close to \(M_{0}'\).

In Fig. 1, we plot the ratio \(M_{0}/M_{0}'\) as a function of the redshift for different values of (w 0,w a ). M 0, as argued, turns to be of order (of a few times) \(M_{0}'\).

Appendix C: Virial overdensity and virial radius

Virial overdensity

The virial overdensity is the ratio of the cluster mass density and the background matter density at the time of virialization:

$$ \varDelta_v(z=z_v) = \frac{\rho_{\mathrm{cluster}}(z_v)}{\rho_m(z_v)} , $$
(C.1)

where z v is the redshift at the time of virialization.

Using the fact that the cluster mass density is

$$ \rho_{\mathrm{cluster}}(z) = \frac{3M_{\mathrm{cluster}}}{4\pi r^3(z)} , $$
(C.2)

where M cluster is the halo cluster mass and r(z) the physical halo radius, we can rewrite Eq. (C.1) as

$$ \varDelta_v(z=z_v) = \biggl( \frac{r_{\mathrm{ta}}}{r_v} \biggr)^{3} \biggl( \frac{1 + z_{\mathrm{ta}}}{1+ z_v} \biggr)^{3} \zeta , $$
(C.3)

where r ta and r v are, respectively, the physical radii of the halo cluster at turn-around and at virialization (r v is, in other words, the physical virial radius of the halo cluster), while z ta and

$$ \zeta = \frac{\rho_{\mathrm{cluster}}(z_{\mathrm{ta}})}{\rho_m(z_{\mathrm{ta}})} $$
(C.4)

are, respectively, the redshift and virial overdensity at the time of turn-around.

The redshift at the virialization time is

$$ z_v = z_{\mathrm{collapse}}, $$
(C.5)

and follows from the standard assumption that clusters virialize at the time of collapse.

To find the redshift at the time of turn-around, z ta, and to obtain the virial overdensity at the turn-around time, ζ, we follow the procedure of Ref. [103]. First, they observed that the quantity (δ+1)/a 3, where δ is the density contrast that satisfies the nonlinear Eq. (A.1), is proportional to 1/r 3(z):

$$ \frac{\delta +1}{a^3} = \frac{3M_{\mathrm{cluster}}}{4\pi \rho_m^{(0)}} \frac{1}{r^3(z)} , $$
(C.6)

where r(z) is the collapsing sphere’s radius. [It is straightforward to get the above equation by using Eqs. (A.2) and (C.2).] Then since r(z) assumes the maximum value at the turn-around time, the time of turn-around is found by minimizing the quantity (δ+1)/a 3, where δ is the solution of Eq. (A.1) obtained by imposing the boundary conditions used in Appendix A.

Once the turn-around time is found, the virial overdensity at the turn-around time is simply given, from Eqs. (A.2) and (C.4), by

$$ \zeta = \delta(z_{\mathrm{ta}}) + 1 , $$
(C.7)

where δ is the solution of Eq. (A.1).

The ratio of the virial radius to the turn-around radius, r v /r ta, as a function of cosmological parameters is analyzed below.

In Fig. 1 we plot the virial overdensity at the time of virialization, Δ v , as a function of the redshift for different values of (w 0,w a ). In an Einstein–de Sitter model (i.e., for Ω m =1), the standard assumption that t(z v )≃2t(z ta) together with the fact that r v =r ta/2 and ζ=(3π/4)2, gives Δ v ≃18π 2≃177.653 (independent of the redshift). Indeed, each curve in Fig. 1 approaches this value for a sufficiently large value of the redshift since, as already noted, the Universe then enters the Einstein–de Sitter regime where the effects of dark energy become subdominant with respect to those of nonrelativistic matter.

In Fig. 8, we show the Δ v -isocontours in the (w 0,w a )-plane for different values of the redshift and for Ω m =0.27.

Fig. 8
figure 8

Δ v -isocontours in the (w 0,w a )-plane for different values of the redshift. We fixed Ω m =0.27

Virial radius

In order to “estimate” the quantity r v /r ta, we apply energy conservation and the virial theorem to the spherical collapse of the cluster halo.

We start by considering the total gravitational potential energy U(r) of a sphere of radius r containing the cluster mass M cluster and dark energy:

$$ U(r,z) = U_{mm}(r) + U_{m \mathrm{DE}}(r,z) , $$
(C.8)

where

$$ U_{mm}(r) = -\frac{3}{5} \frac{GM_{\mathrm{cluster}}^2}{r} $$
(C.9)

is the familiar gravitational potential self-energy of a sphere of nonrelativistic matter, and

$$ U_{m \mathrm{DE}}(r,z) = -\frac{4\pi}{5} GM_{\mathrm{cluster}} \bigl[1+3w(z)\bigr] \rho_{\mathrm{DE}}(z) r^2 $$
(C.10)

is the gravitational potential energy of interaction between nonrelativistic matter and dark energy.Footnote 13

Since the potential energy U mDE(r,z) depends explicitly on the time, the system under consideration is not conservative. Therefore, neither energy conservation nor the virial theorem can be applied.Footnote 14

In order to get a conservative system, Wang [107] has suggested replacing the z-dependent quantity [1+3w(z)]ρ DE(z) with the same quantity evaluated at the turn-around time.

Here we propose defining an effective potential energy, which does not depend explicitly on the time, as

$$ U^{(\mathrm{eff})}(r) = U_{mm}(r) + U^{(\mathrm{eff})}_{m \mathrm{DE}}(r), $$
(C.11)

where

$$ U^{(\mathrm{eff})}_{m \mathrm{DE}}(r) = -\frac{4\pi}{5} GM_{\mathrm{cluster}} \bigl\langle \bigl[1+3w(z)\bigr] \rho_{\mathrm{DE}}(z) \bigr \rangle r^2 $$
(C.12)

is the effective potential energy of interaction and 〈…〉 is an operator that when applied to a z-dependent function ψ(z) gives a z-independent quantity, i.e.

$$ \frac{d \langle \psi(z) \rangle}{dz} = 0 . $$
(C.13)

The action of the 〈…〉-operator is specified below.

The introduction of the effective energy potentials (C.11) and (C.12), allow us to use the energy conservation theorem that, applied at the times of virialization and turn-around, gives

$$ \mathcal{K}(r_v) + U^{(\mathrm{eff})}(r_v) = U^{(\mathrm{eff})}(r_{\mathrm{ta}}) , $$
(C.14)

where \(\mathcal{K}(r_{v})\) is the kinetic energy at the virialization time.

Using the virial theorem

$$ \mathcal{K}(r_v) = \biggl( \frac{r}{2} \frac{dU^{(\mathrm{eff})}(r)}{dr} \biggr)_{r = r_v} , $$
(C.15)

the energy conservation equation (C.14) takes the form

$$ \frac{1}{2} U_{mm}(r_v) + 2 U^{(\mathrm{eff})}_{m \mathrm{DE}}(r_v) = U_{mm}(r_{\mathrm{ta}}) + U^{(\mathrm{eff})}_{m \mathrm{DE}}(r_{\mathrm{ta}}) . $$
(C.16)

Taking into account Eqs. (C.2) and (C.4), the mass of the cluster is

$$ M_{\mathrm{cluster}} = \frac{4\pi}{3} r_{\mathrm{ta}}^3 \rho_m(z_{\mathrm{ta}}) \zeta , $$
(C.17)

so that Eq. (C.16) reads

(C.18)

Here w ta=w(z ta) and we have defined, following [107], the quantity

(C.19)

We have also introduced the “deviation parameter”, η, as

(C.20)

It is worth noting that in the case of a cosmological constant the system is conservative (see footnote 14), and the equation determining r v /r ta is formally given by Eq. (C.18) with η=1. Hence, the only restriction to the action of the 〈…〉-operator is that, when applied to the function [1+3w(z)]ρ DE(z), it must give η=1 for w(z)=−1.

Taking 〈ψ(z)〉=ψ(z ta) corresponds to the choice of Wang, which also implies η=1 for equation-of-state parameter w(z). However, taking \(\langle \psi(z) \rangle = \psi(\bar{z})\), where \(\bar{z}\) can be anywhere in the interval [z v ,z ta], is also a plausible choice.

In the upper panel of Fig. 9, we plot the ratio of the virial radius to the turn-around radius in the case η=1 (Wang’s choice) as a function of the redshift for different values of (w 0,w a ). In the lower panel, we show the same ratio for the case 〈ψ(z)〉=ψ(z v ). For notational clarity, we indicate those ratios by

$$ \frac{r_v}{r_{\mathrm{ta}}} \equiv \left \{ \begin{array}{l@{\quad}l} x & \mbox{if} \ \eta = 1 , \\[6pt] y & \mbox{if} \ \eta = \frac{(1+3w_v) \rho_{\mathrm{DE}}(z_v)}{(1+3w_{\mathrm{ta}}) \rho_{\mathrm{DE}}(z_{\mathrm{ta}})} , \end{array} \right . $$
(C.21)

where w v =w(z v ). For z≫1, where dark energy effects can be neglected, the ratio of the virial to the turn-around radii approaches, as expected, the asymptotic Einstein–de Sitter value x=1/2. The variations of x with the choice of the functional form of the deviation parameter η are of order of a few percent. The resulting analysis on the growth of massive galaxy clusters does not appreciably depend on η, and the results presented in the previous sections are for the case η=1.

Fig. 9
figure 9

Upper panel. The ratio of the virial radius to the turn-around radius, xr v /r ta with η=1 (corresponding to Wang’s choice), as a function of the redshift for different values of (w 0,w a ) (the same as in Fig. 1). Lower panel. The ratio of x to y, where yr v /r ta with η defined in Eq. (C.21). In both panels, we fixed Ω m =0.27

Rights and permissions

Reprints and permissions

About this article

Cite this article

Campanelli, L., Fogli, G.L., Kahniashvili, T. et al. Galaxy cluster number count data constraints on cosmological parameters. Eur. Phys. J. C 72, 2218 (2012). https://doi.org/10.1140/epjc/s10052-012-2218-4

Download citation

  • Received:

  • Revised:

  • Published:

  • DOI: https://doi.org/10.1140/epjc/s10052-012-2218-4

Keywords

Navigation