Skip to main content
Log in

Pattern dark matter and galaxy scaling relations

Is dark matter the self-organized behavior and manifestation of things we already knew?

  • Regular Article
  • Published:
The European Physical Journal Special Topics Aims and scope Submit manuscript

Abstract

We argue that a natural explanation for a variety of robust galaxy scaling relations comes from the perspective of pattern formation and self-organization as a result of symmetry breaking. We propose a simple Lagrangian model that combines a conventional model for normal matter in a galaxy with a conventional model for stripe pattern formation in systems that break continuous translation invariance. We show that the energy stored in the pattern field acts as an effective dark matter. Our theory reproduces the gross features of elliptic galaxies as well as disk galaxies (HSB and LSB) including their detailed rotation curves, the radial acceleration relation, and the Freeman law. We investigate the stability of disk galaxies in the context of our model and obtain scaling relations for the central dispersion for elliptical galaxies. A natural interpretation of our results is that (1) ‘dark matter’ is potentially a collective, emergent phenomenon and not necessarily an as yet undiscovered particle, and (2) MOND is an effective theory for the description of a self-organized complex system rather than a fundamental description of nature that modifies Newton’s second law.

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
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

Data Availability Statement

This manuscript has no associated data or the data will not be deposited. [Authors’ comment: No observational data was generated in writing this paper. The code associated with this paper is publicly available at https://zenodo.org/record/4456019.]

References

  1. V. Trimble, Annu. Rev. Astron. Astrophys. 25, 425 (1987). https://doi.org/10.1146/annurev.aa.25.090187.002233

    Article  ADS  Google Scholar 

  2. J.H. Oort, Bull. Astron. Inst. Neth. 6, 249 (1932)

    ADS  Google Scholar 

  3. F. Zwicky, Helv. Phys. Acta 6, 110 (1933)

    ADS  Google Scholar 

  4. V.C. Rubin, W.K. Ford Jr., Astrophys. J. 159, 379 (1970). https://doi.org/10.1086/150317

    Article  ADS  Google Scholar 

  5. V.C. Rubin, N. Thonnard, W.K. Ford Jr., Astrophys. J. Lett. 225, L107 (1978). https://doi.org/10.1086/182804

    Article  ADS  Google Scholar 

  6. T.S. van Albada et al., Astrophys. J. 295, 305 (1985). https://doi.org/10.1086/163375

    Article  ADS  Google Scholar 

  7. F. Zwicky, Astrophys. J. 86, 217 (1937). https://doi.org/10.1086/143864

    Article  ADS  Google Scholar 

  8. V.C. Rubin, W.K. Ford Jr., N. Thonnard, Astrophys. J. 238, 471 (1980). https://doi.org/10.1086/158003

    Article  ADS  Google Scholar 

  9. M. Milgrom, Astrophys. J. 270, 365 (1983). https://doi.org/10.1086/161130

    Article  ADS  Google Scholar 

  10. J.F. Navarro, C.S. Frenk, S.D.M. White, Astrophys. J. 462, 563 (1996). https://doi.org/10.1086/177173

    Article  ADS  Google Scholar 

  11. P. Li, F. Lelli, S. McGaugh, J. Schombert, Astrophys. J. Suppl. Ser. 247(1), 31 (2020). https://doi.org/10.3847/1538-4365/ab700e

    Article  ADS  Google Scholar 

  12. J. Binney, S. Tremaine, Galactic Dynamics, 2nd edn. (Princeton University Press, Princeton, 2008)

    Book  Google Scholar 

  13. F. Donato, G. Gentile, P. Salucci, Mon. Not. R. Astron. Soc. 353(2), L17 (2004). https://doi.org/10.1111/j.1365-2966.2004.08220.x

    Article  ADS  Google Scholar 

  14. F. Donato, G. Gentile, P. Salucci, C. Frigerio Martins, M.I. Wilkinson, G. Gilmore, E.K. Grebel, A. Koch, R. Wyse, Mon. Not. R. Astron. Soc. 397(3), 1169 (2009). https://doi.org/10.1111/j.1365-2966.2009.15004.x

    Article  ADS  Google Scholar 

  15. S.S. McGaugh, J.M. Schombert, G.D. Bothun, W.J.G. de Blok, Astrophys. J. Lett. 533(2), L99 (2000). https://doi.org/10.1086/312628

    Article  ADS  Google Scholar 

  16. B. Famaey, S.S. McGaugh, Living Rev. Relativ. 15(1), 10 (2012). https://doi.org/10.12942/lrr-2012-10

    Article  ADS  Google Scholar 

  17. S.S. McGaugh, Astron. J. 143(2), 40 (2012). https://doi.org/10.1088/0004-6256/143/2/40

    Article  ADS  Google Scholar 

  18. M.J. Disney, J.D. Romano, D.A. Garcia-Appadoo, A.A. West, J.J. Dalcanton, L. Cortese, Nature 455(7216), 1082 (2008). https://doi.org/10.1038/nature07366

    Article  ADS  Google Scholar 

  19. F. Lelli, S.S. McGaugh, J.M. Schombert, M.S. Pawlowski, Astrophys. J. 836(2), 152 (2017). https://doi.org/10.3847/1538-4357/836/2/152

    Article  ADS  Google Scholar 

  20. M.J. Aschwanden, F. Scholkmann, W. Béthune, W. Schmutz, V. Abramenko, M.C.M. Cheung, D. Müller, A. Benz, G. Chernov, A.G. Kritsuk, J.D. Scargle, A. Melatos, R.V. Wagoner, V. Trimble, W.H. Green, Space Sci. Rev. 214(2), 55 (2018). https://doi.org/10.1007/s11214-018-0489-2

    Article  ADS  Google Scholar 

  21. S.C. Venkataramani, A.C. Newell, Phys. Lett. B 813, 136060 (2021). https://doi.org/10.1016/j.physletb.2020.136060

    Article  Google Scholar 

  22. S.M. Faber, R.E. Jackson, Astrophys. J. 204, 668 (1976). https://doi.org/10.1086/154215

    Article  ADS  Google Scholar 

  23. S. Djorgovski, M. Davis, Astrophys. J. 313, 59 (1987). https://doi.org/10.1086/164948

    Article  ADS  Google Scholar 

  24. D.H. Gudehus, Astrophys. J. 382, 1 (1991). https://doi.org/10.1086/170687

    Article  ADS  Google Scholar 

  25. Y. Couder, E. Fort, Phys. Rev. Lett. 97, 154101 (2006). https://doi.org/10.1103/PhysRevLett.97.154101

    Article  ADS  Google Scholar 

  26. J.W.M. Bush, Proc. Natl. Acad. Sci. 107(41), 17455 (2010). https://doi.org/10.1073/pnas.1012399107. https://www.pnas.org/content/107/41/17455

  27. J. Moláček, J.W.M. Bush, J. Fluid Mech. 727, 582 (2013). https://doi.org/10.1017/jfm.2013.279. https://www.cambridge.org/core/article/drops-bouncing-on-a-vibrating-bath/441A614F657E800EA06F6C08674CCE67

  28. L.D. Landau, Zh. Eksp. Teor. Fiz. 7, 19 (1937)

    Google Scholar 

  29. L.D. Landau, Ukr. J. Phys. 53, 25 (2008)

    Google Scholar 

  30. T. Passot, A.C. Newell, Phys. D Nonlinear Phenom. 74(3–4), 301 (1994). https://doi.org/10.1016/0167-2789(94)90199-6

    Article  ADS  Google Scholar 

  31. A.C. Newell, S.C. Venkataramani, Stud. Appl. Math. 139(2), 322 (2017). https://doi.org/10.1111/sapm.12184

    Article  MathSciNet  Google Scholar 

  32. A.C. Newell, S.C. Venkataramani, Comptes Rendus Mécanique 347(4), 318 (2019). https://doi.org/10.1016/j.crme.2019.03.004

    Article  ADS  Google Scholar 

  33. A.C. Newell, T. Passot, C. Bowman, N. Ercolani, R. Indik, Phys. D Nonlinear Phenom. 97(1), 185 (1996)

    Article  ADS  Google Scholar 

  34. C.W. Misner, K.S. Thorne, J.A. Wheeler, Gravitation (W.H. Freeman and Co., San Francisco, 1973)

    Google Scholar 

  35. T.A. Driscoll, N. Hale, L.N. Trefethen, Chebfun Guide (Pafnuty Publications, 2014). http://www.chebfun.org/docs/guide/

  36. R. Sancisi, Symp. IAU 220, 233 (2004). https://doi.org/10.1017/S0074180900183299

    Article  ADS  Google Scholar 

  37. B. Famaey, J. Khoury, R. Penco, A. Sharma, J. Cosmol. Astropart. Phys. 2020(06), 025 (2020). https://doi.org/10.1088/1475-7516/2020/06/025

    Article  Google Scholar 

  38. S. Boyd, L. Vandenberghe, Convex Optimization (Cambridge University Press, Cambridge, 2004)

    Book  Google Scholar 

  39. J.P. Ostriker, P.J.E. Peebles, Astrophys. J. 186, 467 (1973). https://doi.org/10.1086/152513

    Article  ADS  Google Scholar 

  40. M. Milgrom, Astrophys. J. 338, 121 (1989). https://doi.org/10.1086/167184

    Article  ADS  Google Scholar 

  41. R. Brada, M. Milgrom, Astrophys. J. 519(2), 590 (1999). https://doi.org/10.1086/307402

    Article  ADS  Google Scholar 

  42. G. Bertin, Dynamics of Galaxies (Cambridge University Press, Cambridge, 2014)

    Book  Google Scholar 

  43. A.B. Romeo, K.M. Mogotsi, Mon. Not. R. Astron. Soc. 480(1), L23 (2018). https://doi.org/10.1093/mnrasl/sly119

    Article  ADS  Google Scholar 

  44. A.B. Romeo, Mon. Not. R. Astron. Soc. 491(4), 4843 (2020). https://doi.org/10.1093/mnras/stz3367

    Article  ADS  Google Scholar 

  45. A.B. Romeo, O. Agertz, F. Renaud, Mon. Not. R. Astron. Soc. 499(4), 5656 (2020). https://doi.org/10.1093/mnras/staa3245

    Article  ADS  Google Scholar 

  46. J. Rutter, Geometry of Curves (Chapman & Hall/CRC, Boca Raton, 2000)

    MATH  Google Scholar 

  47. S.S. McGaugh, G.D. Bothun, J.M. Schombert, Astron. J. 110, 573 (1995). https://doi.org/10.1086/117543

    Article  ADS  Google Scholar 

  48. M. Abramowitz, I.A. Stegun (eds.), Handbook of mathematical functions with formulas, graphs, and mathematical tables (Dover Publications Inc., New York, 1992). Reprint of the 1972 edition

  49. K.C. Freeman, Astrophys. J. 160, 811 (1970). https://doi.org/10.1086/150474

    Article  ADS  Google Scholar 

  50. C. Di Paolo, P. Salucci, A. Erkurt, Mon. Not. R. Astron. Soc. 490(4), 5451 (2019). https://doi.org/10.1093/mnras/stz2700

    Article  ADS  Google Scholar 

  51. S.W. Morris, E. Bodenschatz, D.S. Cannell, G. Ahlers, Phys. D Nonlinear Phenom. 97(1), 164 (1996)

    Article  ADS  Google Scholar 

  52. N. Kopell, L.N. Howard, Adv. Appl. Math. 2(4), 417 (1981)

    Article  MathSciNet  Google Scholar 

  53. G. de Vaucouleurs, Annales d’Astrophysique 11, 247 (1948)

    ADS  Google Scholar 

  54. W. Jaffe, Mon. Not. R. Astron. Soc. 202, 995 (1983). https://doi.org/10.1093/mnras/202.4.995

    Article  ADS  Google Scholar 

  55. V.F. Cardone, G. Angus, A. Diaferio, C. Tortora, R. Molinaro, Mon. Not. R. Astron. Soc. 412(4), 2617 (2011). https://doi.org/10.1111/j.1365-2966.2010.18081.x

    Article  ADS  Google Scholar 

  56. R. Brada, M. Milgrom, Mon. Not. R. Astron. Soc. 276(2), 453 (1995). https://doi.org/10.1093/mnras/276.2.453

    Article  ADS  Google Scholar 

  57. J. Bekenstein, M. Milgrom, Astrophys. J. 286, 7 (1984). https://doi.org/10.1086/162570

    Article  ADS  Google Scholar 

  58. J.D. Bekenstein, Phys. Rev. D 70, 083509 (2004). https://doi.org/10.1103/PhysRevD.70.083509

    Article  ADS  Google Scholar 

  59. C. Skordis, T. Złosnik, A new relativistic theory for Modified Newtonian Dynamics (2020). ArXiv e-prints

  60. S.S. McGaugh, F. Lelli, J.M. Schombert, Phys. Rev. Lett. 117, 201101 (2016). https://doi.org/10.1103/PhysRevLett.117.201101

    Article  ADS  Google Scholar 

  61. W. Dehnen, Astron. J. 118(3), 1190 (1999). https://doi.org/10.1086/301009

    Article  ADS  Google Scholar 

  62. W. Dehnen, Astron. J. 118(3), 1201 (1999). https://doi.org/10.1086/301010

    Article  ADS  Google Scholar 

  63. C. Di Paolo, P. Salucci, J.P. Fontaine, Astrophys. J. 873(2), 106 (2019). https://doi.org/10.3847/1538-4357/aaffd6

    Article  ADS  Google Scholar 

  64. S.J. Sin, Phys. Rev. D 50, 3650 (1994). https://doi.org/10.1103/PhysRevD.50.3650

    Article  ADS  Google Scholar 

  65. D.N. Spergel, P.J. Steinhardt, Phys. Rev. Lett. 84, 3760 (2000). https://doi.org/10.1103/PhysRevLett.84.3760

    Article  ADS  Google Scholar 

  66. L. Berezhiani, J. Khoury, Phys. Rev. D 92, 103510 (2015). https://doi.org/10.1103/PhysRevD.92.103510

    Article  ADS  Google Scholar 

  67. W. Hu, R. Barkana, A. Gruzinov, Phys. Rev. Lett. 85, 1158 (2000). https://doi.org/10.1103/PhysRevLett.85.1158

    Article  ADS  Google Scholar 

  68. A. Toomre, Astrophys. J. 139, 1217 (1964). https://doi.org/10.1086/147861

    Article  ADS  Google Scholar 

  69. C. Deffayet, G. Esposito-Farèse, R.P. Woodard, Phys. Rev. D 84, 124054 (2011). https://doi.org/10.1103/PhysRevD.84.124054

    Article  ADS  Google Scholar 

  70. N.M. Ercolani, S.C. Venkataramani, J. Nonlinear Sci. 19(3), 267 (2009). https://doi.org/10.1007/s00332-008-9035-9

    Article  ADS  MathSciNet  Google Scholar 

  71. P. Hořava, Phys. Rev. D 79, 084008 (2009). https://doi.org/10.1103/PhysRevD.79.084008

    Article  ADS  MathSciNet  Google Scholar 

  72. D. Blas, O. Pujolàs, S. Sibiryakov, Phys. Rev. Lett. 104, 181302 (2010). https://doi.org/10.1103/PhysRevLett.104.181302

    Article  ADS  MathSciNet  Google Scholar 

  73. J.J. Dalcanton, D.N. Spergel, F.J. Summers, Astrophys. J. 482(2), 659 (1997). https://doi.org/10.1086/304182

    Article  ADS  Google Scholar 

  74. R.H. Wechsler, J.L. Tinker, Ann. Rev. Astron. Astrophys. 56(1), 435 (2018). https://doi.org/10.1146/annurev-astro-081817-051756

    Article  ADS  Google Scholar 

  75. M. Milgrom, Astrophys. J. 698(2), 1630 (2009). https://doi.org/10.1088/0004-637x/698/2/1630

    Article  ADS  Google Scholar 

  76. L.P. Kadanoff, Phys. A Stat. Mech. Appl. 163(1), 1 (1990). https://doi.org/10.1016/0378-4371(90)90309-G

    Article  Google Scholar 

  77. J.F. Navarro, A. Benítez-Llambay, A. Fattahi, C.S. Frenk, A.D. Ludlow, K.A. Oman, M. Schaller, T. Theuns, Mon. Not. R. Astron. Soc. 471(2), 1841 (2017). https://doi.org/10.1093/mnras/stx1705

    Article  ADS  Google Scholar 

  78. A.A. Dutton, A.V. Macciò, A. Obreja, T. Buck, Mon. Not. R. Astron. Soc. 485(2), 1886 (2019). https://doi.org/10.1093/mnras/stz531

  79. E.P. Verlinde, SciPost Phys. 2, 016 (2017). https://doi.org/10.21468/SciPostPhys.2.3.016

  80. S. Nandi, Am. J. Phys. 78(12), 1341 (2010). https://doi.org/10.1119/1.3474665

    Article  ADS  Google Scholar 

  81. D.J. Griffiths, D.F. Schroeter, Introduction to Quantum Mechanics (Cambridge University Press, Cambridge, 2018)

    Book  Google Scholar 

Download references

Acknowledgements

SCV was partially supported by the Simons Foundation through award 524875. SCV and ACN were also partially supported by the National Science Foundation through award GCR-2020915.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shankar C. Venkataramani.

Appendices

Appendix A: The stress tensor for pattern dark matter

We now follows the discussion in [32] and compute the stress–energy–momentum \(T^{\alpha \beta }\) corresponding to this solution using the Einstein–Hilbert prescription \(T^{\alpha \beta } = \frac{2}{\sqrt{-g}} \frac{\delta {\tilde{E}}}{\delta g_{\alpha \beta }}\) where \({\tilde{E}}\) is the appropriate Lagrangian in curved space-time and \(g = \mathrm {det}[g_{\alpha \beta }]\). The minimally coupled [34] pattern action \({\mathcal {S}}_\mathrm{{P}}\) is:

$$\begin{aligned} {\mathcal {S}}_\mathrm{{P}} = \frac{{\varSigma }^*_0 c^2}{k_0^3} \int \left\{ (\nabla ^\mu \psi \nabla _\mu \psi -k_0^2)^2 + (\nabla ^\mu \nabla _\mu \psi )^2\right\} \sqrt{-g} \ d^4x, \nonumber \\ \end{aligned}$$
(A.1)

where the metric \(g_{\alpha \beta }\) has signature \((- \,+\, +\, +)\), \(\nabla _\mu \) is the corresponding covariant derivative. To obtain the stress tensor in a (background) flat space time, it suffices to consider variations \(g_{\alpha \beta } = \eta _{\alpha \beta } + t \rho _{\alpha \beta }\) and compute all the variations to first order in t. To this end, we record the following relations for the inverse metric \(g^{\gamma \delta }\), the Christoffel symbols \({\varGamma }^\sigma _{\gamma \delta }\) and the quantities that appear in \({\mathcal {S}}_\mathrm{{P}}\):

$$\begin{aligned} g^{\gamma \delta }&= \eta ^{\gamma \delta } - t \, \eta ^{\gamma \alpha }\eta ^{\delta \beta } \rho _{\alpha \beta } + O(t^2) \\ {\varGamma }^\sigma _{\gamma \delta }&= \frac{t}{2}\, \eta ^{\sigma \xi } \left[ \frac{\partial \rho _{\gamma \xi }}{\partial x^\delta } + \frac{\partial \rho _{\delta \xi }}{\partial x^\gamma } - \frac{\partial \rho _{\gamma \delta }}{\partial x^\xi }\right] + O(t^2) \\ \sqrt{-g}&= 1 + \frac{t}{2} \eta ^{\alpha \beta } \rho _{\alpha \beta } + O(t^2) \\ \nabla ^\mu \psi \nabla _\mu \psi&= \eta ^{\gamma \delta } \partial _\gamma \psi \partial _\delta \psi - t \, \eta ^{\gamma \alpha }\eta ^{\delta \beta } \rho _{\alpha \beta } \partial _\gamma \psi \partial _\delta \psi + O(t^2) \\ g^{\gamma \delta } \nabla _\gamma \nabla _\delta \psi&= \Box \, \psi - t \, \eta ^{\gamma \alpha }\eta ^{\delta \beta } \rho _{\alpha \beta } \partial _{\gamma } \partial _{\delta } \psi \\&\quad - \frac{t}{2}\, \eta ^{\gamma \delta } \eta ^{\sigma \xi } \left[ \frac{\partial \rho _{\gamma \xi }}{\partial x^\delta } + \frac{\partial \rho _{\delta \xi }}{\partial x^\gamma } - \frac{\partial \rho _{\gamma \delta }}{\partial x^\xi }\right] \partial _\sigma \psi + O(t^2) \end{aligned}$$

The Einstein–Hilbert stress tensor is now given by

$$\begin{aligned} \left. \frac{\mathrm{{d}}}{\mathrm{{d}}t}{\mathcal {S}}_\mathrm{{P}}\right| _{t=0} = \frac{1}{2} \int T^{\alpha \beta } \rho _{\alpha \beta } \,d^4x + \hbox { boundary terms}. \end{aligned}$$

A straightforward but somewhat lengthy calculation now yields

$$\begin{aligned} T^{\alpha \beta }&= \frac{{\varSigma }^* c^2}{k_0^3}\left\{ -4(\eta ^{\mu \nu } \partial _\mu \psi \partial _\nu \psi -k_0^2) \eta ^{\sigma \alpha }\eta ^{\tau \beta } \partial _\sigma \psi \partial _\tau \psi \right. \nonumber \\&\quad \left. - 4\, \Box \psi \, \eta ^{\sigma \alpha }\eta ^{\tau \beta } \partial _{\sigma } \partial _{\tau } \psi \right. \nonumber \\&\quad + 2 (\eta ^{\alpha \tau } \eta ^{ \beta \sigma } + \eta ^{\beta \tau } \eta ^{\alpha \sigma } - \eta ^{\alpha \beta } \eta ^{\sigma \tau } ) \partial _\tau (\Box \psi \, \partial _\sigma \psi ) \nonumber \\&\quad + \left. \eta ^{\alpha \beta } \left[ (\eta ^{\mu \nu } \partial _\mu \psi \partial _\nu \psi -k_0^2)^2 + (\Box \psi )^2\right] \right\} \end{aligned}$$
(A.2)

We can now express the energy density \(T^{\alpha \beta }\) for the (stationary) phase field \(\psi (R)\) with respect to a normalized basis \(\{{\mathbf {e}}_t, {\mathbf {e}}_R, {\mathbf {e}}_\theta , {\mathbf {e}}_\phi \}\) induced by (spatial) spherical polar coordinates \((r,\theta ,\phi )\).

We decompose the stress tensor into two pieces, \(T_\mathrm{{s}}\) coming from the “stretching energy” with density \((|\nabla \psi |^2 - c^{-2} (\partial _t \psi )^2 - k_0^2))^{2}\) and \(T_\mathrm{{b}}\) from the bending energy \((\Box \, \psi )^2\). We compute these quantities for a stationary, radial solution \(\psi = \psi (r)\) to get

$$\begin{aligned}&T_\mathrm{{s}} \sim \frac{{\varSigma }^*c^2}{k_0^3} \begin{pmatrix} \tau _1 &{} 0 &{} 0 &{} 0 \\ 0 &{} \tau _2 &{} 0 &{} 0 \\ 0 &{} 0 &{}-\tau _1 &{} 0 \\ 0 &{} 0 &{} 0 &{} -\tau _1 \end{pmatrix} \nonumber \\&T_\mathrm{{b}} \sim \frac{{\varSigma }^*c^2}{k_0^3} \begin{pmatrix} \tau _3 &{} 0 &{} 0 &{} 0 \\ 0 &{} \tau _4 &{} 0 &{} 0 \\ 0 &{} 0 &{}-\tau _3 &{} 0 \\ 0 &{} 0 &{} 0 &{} -\tau _3 \end{pmatrix} \end{aligned}$$
(A.3)

where

$$\begin{aligned} \tau _1&= (\psi '(R)^2-k_0^2)^2 \nonumber \\ \tau _2&= (3 \psi '(R)^2 +k_0^2)(\psi '(R)^2- k_0^2) \nonumber \\ \tau _3&= -\psi ''(R)^2-\frac{2 \psi '(R) \left( R \psi '''(R)+4 \psi ''(R)\right) }{R} \nonumber \\ \tau _4&= \frac{8 \psi '(R)^2}{R^2}+\psi ''(R)^2-2 \psi '''(R) \psi '(R) \end{aligned}$$
(A.4)

We remark on the expected structure of \(T_\mathrm{{s}}\) and \(T_\mathrm{{b}}\), viz. the off-diagonal stresses should be zero from time-reversal and spherical symmetries of the solution \(\psi \), and further \(T^{00} = -T^{\theta \theta } = - T^{\phi \phi }\) since the metric has signature \((- \,+\, +\, +)\), and \(\psi \) is independent of \(t, \theta \) and \(\phi \). Finally, the quantities \(\tau _1, \tau _2, \tau _3\) and \(\tau _4\) are constrained by the conservation of energy–momentum \(\nabla _\mu T^{\mu \nu } = 0\). A calculation shows that, as expected, these 4 conditions reduce to just one constraint on \(\psi (r)\), namely that \(\psi \) should satisfy the Euler–Lagrange equation (4).

The stress tensor associated with the field \(\psi \) will act as source for the curvature of space-time as we discuss in the body of this paper.

Appendix B: Bound states for Gaussian wells

Following the discussion in Sect. 5.1, we now estimate the ground state energies for spherical and oblate Gaussian wells, to determine the appropriate choices for \(k_0\) for spherical (i.e. elliptic) and disk galaxies.

The 1d Gaussian well is given by a potential \(V(x) = - V_0 e^{-\frac{x^2}{2 a^2}}\). Close to the minimum, this is approximated by the Harmonic potential \(-V_0 + \frac{V_0 x^2}{2 a^2}\) suggesting that an estimate can be obtained by using the variational method with trial functions given by eigenfunctions for the Harmonic oscillator [80]. With the normalized wavefunction \(\psi _0 = \left( \frac{\beta }{\pi }\right) ^{1/4} e^{-\beta x^2/2}\), we have, from the variational principle [81],

$$\begin{aligned} \int \left( \psi _0'^2 + V(x) \psi _0^2 \right) dx = \frac{\beta }{2} - V_0 \sqrt{\frac{2 a^2 \beta }{1+2a^2\beta }} \ge -k_0^2 \end{aligned}$$

where \(-k_0^2\) is the ground state energy for the operator \(-{\varDelta } + V\). For small \(\beta > 0\), the bound is \(\frac{\beta }{2} - V_0 a \sqrt{2 \beta }\) which can be made negative independent of how small \(V_0\) is, i.e. we always have bound states. The optimal \(\beta \) minimizes the bound, and is given by

$$\begin{aligned} \frac{1}{2}= & {} V_0 \sqrt{\frac{2 a^2 \beta }{1+2a^2\beta }} \left[ \frac{1}{\beta } - \frac{2a^2}{1+2a^2\beta } \right] \\= & {} a^2 V_0 \sqrt{\frac{2}{a^2\beta (1+2 a^2 \beta )^3}} \end{aligned}$$

It is clear that the optimal \(\beta \) is in the “scaling form” \(\beta = a^{-2} F(a^2 V_0)\). Considering the regimes \(a^2 V_0 \gg 1\) and \(a^2 V_0 \ll 1\) separately we get

$$\begin{aligned} a^2 \beta \approx {\left\{ \begin{array}{ll} 8(a^2V_0)^2 &{} a^2 V_0 \ll 1 \\ \sqrt{a^2V_0} &{} a^2 V_0 \gg 1 \end{array}\right. } \end{aligned}$$

In the regime \(a^2 V_0 \gg 1\) we get the “deep potential” limit

$$\begin{aligned} k_0^2 > rsim V_0 - \frac{\sqrt{V_0}}{2a}, \end{aligned}$$
(B.1)

corresponding to the Harmonic oscillator \(V(x) = -V_0 + \frac{V_0}{2a} x^2\). In the complementary regime \(a^2 V_0 \ll 1\), we get cancellation at leading order and

$$\begin{aligned} k_0^2 > rsim 0 (aV_0)^2 + 32 a^6 V_0^4 + \cdots \end{aligned}$$
(B.2)

To investigate if this cancellation is specific to the form of the variational test function that was used, we also consider the normalized wavefunction \(\psi _1 = \left( \beta ' \right) ^{1/4} e^{-\sqrt{\beta '}|x| }\), where the variational parameter \(\beta '\) is chosen so that it is dimensionally consistent with the earlier choice in the definition of \(\psi _0\). The variational principle gives

$$\begin{aligned}&\int \left( \psi _1'^2 + V(x) \psi _1^2 \right) dx\\&\quad =\beta ' - V_0 \,\sqrt{2 \pi a^2 \beta ' } \, \hbox {erfc}\left( \sqrt{2 a^2 \beta '} \right) e^{2 a^2 \beta ' } \ge -k_0^2 \end{aligned}$$

It is again clear that the optimal \(\beta '\) is in the “scaling form” \(\beta ' = a^{-2} F(a^2 V_0)\). Indeed this is immediate from dimensional considerations. Considering the regimes \(a^2 V_0 \gg 1\) and \(a^2 V_0 \ll 1\) separately we get

$$\begin{aligned} a^2 \beta ' \approx {\left\{ \begin{array}{ll} \frac{\pi }{2}(a^2V_0)^2 &{} a^2 V_0 \ll 1 \\ \frac{1}{2}\sqrt{a^2V_0} &{} a^2 V_0 \gg 1 \end{array}\right. } \end{aligned}$$

In the regime \(a^2 V_0 \gg 1\) we get the variational bound

$$\begin{aligned} k_0^2 > rsim V_0 - \frac{\sqrt{V_0}}{a}, \end{aligned}$$
(B.3)

which is consistent in terms of the scaling of the first correction to the leading order behavior, but suboptimal, in comparison with the bound in (B.1). In the complementary regime \(a^2 V_0 \ll 1\), we get

$$\begin{aligned} k_0^2 > rsim \frac{\pi }{2} (aV_0)^2 \end{aligned}$$
(B.4)

showing no cancellation at leading order and giving a result consistent with the limit \(V(x) \rightarrow -\sqrt{2 \pi } aV_0 \delta (x)\).

The 3d isotropic Gaussian well is given by the potential \(V(x_1,x_2,x_3) = -V_0 \exp \left( -\frac{x_1^2+x_2^2 +x_3^2}{2a^2}\right) \). Considering the product test function \({\varPsi }_0(x_1,x_2,x_3) = \psi _0(x_1) \psi _0(x_2) \psi _0(x_3)\) we get the variational bound

$$\begin{aligned} \int \left( |\nabla {\varPsi }_0|^2 + V(x) {\varPsi }_0^2 \right) d^3x = \frac{3\beta }{2} - V_0 \left( \frac{2 a^2 \beta }{1+2a^2\beta }\right) ^{\frac{3}{2}} \ge -k_0^2 \end{aligned}$$

For sufficiently small \(\beta \), the bound is asymptotically given by \(\frac{3 \beta }{2} - \sqrt{8} V_0 a^3 \beta ^{3/2}\), and is positive. Likewise, for large \(\beta \), the bound is asymptotically equal to \(\frac{3 \beta }{2} - V_0\) which is also positive. This reflects the well known fact that attractive potentials in 3d do not support bound states, unless the potential is sufficiently deep.

To estimate the critical value of \(V_0\) that allows for a bound state, we exploit the well known connection between the 3d and 1d Schrödinger equations [81], that follows from the identity

$$\begin{aligned} \frac{\partial ^2 f}{\partial R^2} + \frac{2}{R} \frac{\partial f}{\partial R} = \frac{1}{R} \frac{\partial ^2}{\partial R^2} (R f), \end{aligned}$$

so that for every spherically symmetric eigenfunction \({\varPsi }(R)\) of the 3d operator \(-{\varDelta } + V(R)\) with a spherically symmetric potential, the function \(\chi (R) = R {\varPsi }(R)\) is an eigenfunction of the 1d operator \(-\partial _{RR} + V(R)\) with the same energy.

The converse, however, is not true. In order for the kinetic energy \( 4 \pi \int |\nabla {\varPsi }|^2 R^2 \mathrm{{d}}R\) to be finite, we cannot have \({\varPsi }(R)\) diverging as 1/R near \(R = 0\). We therefore need that the corresponding 1d wavefunction \(R {\varPsi }(R)\) vanish at \(R = 0\), i.e. we need a 1d bound state with a node at the origin to have a bound (ground) state for the 3d potential. The 1st excited state for the 1d gaussian well, which is the lowest energy state with a node at \(R=0\), gives the energy of the ground state for the 3d gaussian well.

Fig. 12
figure 12

The solid curve is energy level \(\lambda _1\) of the first excited state of the 1d Gaussian potential \(V(R) = \eta e^{-R^2/2}\). This state exists only if the well is sufficiently deep \(\eta \ge \eta ^* \approx 1.3\). A quadratic fit (the dashed curve) shows that \(\lambda _1 \approx 0.154(\eta -1.3)^2\)

Figure 12 shows the numerically computed energy [35] of the first excited state for the potential \(W(R) = \eta e^{-R^2/2}\). There is a critical value \(\eta ^* \approx 1.3\) and an O(1) constant \(c \approx 0.154\) such that the first excited energy level satisfies

$$\begin{aligned} \lambda _1 = a^2 k_0^2 \approx c (V_0 a^2 - \eta ^*)^2 . \end{aligned}$$
(B.5)

For applications to LSB disk galaxies, we also consider the anisotropic (oblate) potential \(V(x_1,x_2,x_3) = -V_0 \exp \left( -\frac{x_1^2+x_2^2}{2a^2} - \frac{x_3^2}{2b^2}\right) \) with \(b \ll a\). We now consider the product test function \({\varPsi } =\psi _0(x_1) \psi _0(x_2) \psi _1(x_3)\) to allow for the possibility that \(V_0 a^2 \gg 1\) but \(V_0 b^2 \ll 1\). We now get the variational bound

$$\begin{aligned}&\int \left( |\nabla {\varPsi }|^2 + V(x) {\varPsi }^2 \right) d^3x \\&\quad = \beta +\beta ' - \sqrt{2 \pi b^2 \beta ' } \, \hbox {erfc}\left( \sqrt{2 b^2 \beta '} \right) e^{2 b^2 \beta ' } \frac{2 a^2 \beta V_0}{1+2a^2\beta }\\&\quad \ge -k_0^2 \end{aligned}$$

Since \(b^2 V_0 \ll 1\) it is suggestive that \(\beta ' b^2 \ll 1\). We will assume this provisionally, and verify later that the assumption is valid. With this assumption, we have the variational bound

$$\begin{aligned} k_0^2 \ge \frac{2 a^2 b \beta \sqrt{2 \pi \beta '} V_0}{1+2a^2\beta } - \beta - \beta ' \end{aligned}$$

Optimizing over \(\beta '\), we get that the optimal \(\beta '\) is given by

$$\begin{aligned} b^2 \beta ' = \frac{\pi }{2} \left( \frac{2 a^2 b \beta }{1+2a^2\beta }\right) ^2 (b^2 V_0)^2 \end{aligned}$$

so that our assumption that \(b^2 \beta ' \ll 1\) verifies. Also, we have the variational bound

$$\begin{aligned} k_0^2 \ge \beta \left( \pi \frac{2 a^2 \beta }{(1+2a^2\beta )^2} \cdot (a^2 V_0) \cdot (b^2 V_0) -1\right) \end{aligned}$$
(B.6)

From this bound, it follows that there exists a bound \(\eta _\mathrm{{c}}'\) such that we have bound states for all \(a b V_0 \ge \eta _\mathrm{{c}}'\). The optimal \(\beta \sim a^{-2}\) and for \(ab V_0 \gg 1\), we have \(k_0 \sim b V_0\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Venkataramani, S.C., Newell, A.C. Pattern dark matter and galaxy scaling relations. Eur. Phys. J. Spec. Top. 230, 2139–2165 (2021). https://doi.org/10.1140/epjs/s11734-021-00196-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1140/epjs/s11734-021-00196-9

Navigation