Skip to main content

The Unexpected Resurgence of Weyl Geometry in late 20th-Century Physics

  • Chapter
  • First Online:
Beyond Einstein

Part of the book series: Einstein Studies ((EINSTEIN,volume 14))

Abstract

Weyl’s original scale geometry of 1918 (“purely infinitesimal geometry”) was withdrawn by its author from physical theorizing in the early 1920s. It made a surprising comeback, however, in the last third of the 20th century in several different contexts: scalar-tensor theories of gravity, foundations of space-time theories, foundations of quantum mechanics, elementary particle physics, and cosmology. It seems that Weyl geometry continues to offer an open window for research on the foundations of physics even after the turn into the new millenium.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Such an attempt seemed to be supported experimentally by the phenomenon of (Bjorken) scaling in deep inelastic electron-proton scattering experiments. The latter indicated, at first glance, an active scaling symmetry of mass/energy in high energy physics; but it turned out to hold only approximatively and was of restricted range.

  2. 2.

    SM fields are here called pseudo-classical if they are considered before, or better abstracting from (so-called “second”) quantization. Mathematically they are classical fields (spinor fields or gauge connections), but the field components do not correspond to physically measurable quantities. Observationally relevant information can be extracted only after applying perturbative quantization methods or, in semiclassical approximations, from their wave-function-like probability currents.

  3. 3.

    This distinguishes Euclidean geometry from the other classical geometries of constant curvature. The consideration of material systems, like hydrogen or caesium atoms etc., may be used to introduce units and reduce the allowed automorphisms of the congruences.

  4. 4.

    In mathematical terminology, the implementation of a similarity structure happens at the infinitesimal level. In the following the physicists’ use of the terminology “local” will be used. A discussion, given by Weyl later in his life, of the role of mathematical and physical autormorphisms can be found in (Weyl 1949/2016), some aspects of this also appear in (Weyl 1949, chap. III, sec. 14).

  5. 5.

    For more historical and philosophical details see, among others, (Vizgin 1994; Goenner 2004; Ryckman 2005; Scholz 1999; Scholz et al. 2001).

  6. 6.

    For a modern presentation of Cartan geometry, including the Cartan-Weyl case, see, e.g., (Sharpe 1997, chap. 7); for the physical aspects of the extension studied since the 1970s (Blagojević/Hehl 2013, chap. 8).

  7. 7.

    Presentations of Weyl geometry can be found, among others, in (Blagojević 2002; Israelit 1999b), (Tonnelat 1965, chap. IX), (Drechsler/Tann 1999, appendix A) and (Perlick 1989) (difficult to access). For selected aspects see (Codello et al. 2013) and (Ohanian 2016, sec. 4). Integrable Weyl geometry is presented in (Dahia et al. 2008; Romero et al. 2011; Almeida et al. 2014; Quiros 2014b), (Scholz 2011a, sec. 2.1). Be aware of different conventions for the scale connection. Expressions for Weyl-geometric derivatives and curvature quantities are derived in (Gilkey et al. 2011; Yuan/Huang 2013) and (Miritzis 2004, App.). For a more mathematical perspective consult (Folland 1970; Calderbank/Pedersen 1998; Gauduchon 1995; Higa 1993; Ornea 2001; Gilkey et al. 2011).

  8. 8.

    As the “most simple and natural” expression α 2 = 0 in (Weyl 1918a) and α 1 = 0 as the simplest example in (Weyl 1918b, 4th ed., 5th ed.).

  9. 9.

    In a way, this may be called a geometrical “tensor-vector scalar” theory sui generis, in which all components have geometrical meaning.

  10. 10.

    The dynamical consequences of this interdependence have been clarified by (Israelit 1999b,a), see Section 11.6.2.1.

  11. 11.

    Obviously the Einstein gauge exists also in the non-integrable case.

  12. 12.

    Cf. Sections 11.6.2.2, 11.6.4.

  13. 13.

    A similar approach had been proposed by W. Scherrer in the 1930s but had received no attention by other scientists (Goenner 2012). For recent surveys on scalar tensor theories including JBD gravity see (Capozziello/Faraoni 2011, chap. 3), (Clifton/Ferreira et al. 2012, chap. 3.1)

  14. 14.

    Warning: One has to check carefully the sign convention used in the definition of Riemann and scalar curvature. Jordan, e.g., used sign inverted definitions of the curvature terms with respect to those used here and in much of the present literature (Jordan 1952, 40). In Fujii/Maeda’s notation (see below) this would correspond to 𝜖 = −1 and thus to a “ghost” field.

  15. 15.

    For Dirac’s role in this story see (Kragh 2016), and for a larger view of JBD theory see (Brans 1999, 2014).

  16. 16.

    The conformal factor Ω was (unnecessarily) restricted by the condition Ω 2 = χ γ for some constant γ ∈R.

  17. 17.

    In the 1950s Sciama had considered the possibility that the gravitational “constant” was related to the mass and the “radius” of the visible universe.

  18. 18.

    Also in the English edition of Philosophy of Mathematics and Natural Sciences Weyl expressed this disassociation quite clearly, appealing to the constants of atomic physics which regulate the frequencies of spectral lines (Weyl 1949, 83). But this was only one part of his perspective. In the appendix he argued that for a deeper insight it would be necessary to understand how the “adaptation” of the mass of the electron to the local field constellation is achieved (Weyl 1949, 288f). This was close to the intentions of his 1918 approach, although no longer a claim that the goal had been achieved. Einstein, in his later papers, agreed (Einstein 1949, 555f); see (Lehmkuhl 2014).

  19. 19.

    𝜖 = 1 corresponds to a normal field having a positive energy, in other words, not a “ghost”. Fuji/Maeda add that 𝜖 = −1 looks unacceptable because it seems to indicate negative energy, but “this need not be an immediate difficulty owing to the presence of the nonminimal coupling” (ibid.).

  20. 20.

    See, e.g., (Capozziello/Faraoni 2011, chap. 3.6).

  21. 21.

    See, e.g., (Gilkey et al. 2011).

  22. 22.

    See the discussion in (Quiros et al. 2013) and (Scholz 2017).

  23. 23.

    (Weyl 1920)

  24. 24.

    Weyl’s note (Weyl 1921) became better known by his calculation and discussion of projective and conformal curvature tensors, which followed.

  25. 25.

    See (Trautman 2012).

  26. 26.

    I thank S. Walter for the hint.

  27. 27.

    Among them in particular (Audretsch/Lämmerzahl 1988; Lämmerzahl 1990; Audretsch/Lämmerzahl 1991; Audretsch/Hehl/Lämmerzahl 1992) and (Audretsch/Lämmerzahl 1994).

  28. 28.

    (Audretsch/Hehl/Lämmerzahl 1992, pp. 390ff), compare (Audretsch/Lämmerzahl 1988, eqn. (4.5)).

  29. 29.

    The flow could be characterized by the 0-th order probability current \(j^{\mu }_o\) of the pseudo-classical field. In their argument that the current is geodesic the authors neglected terms proportional to \(j^{\mu }_o\) (Audretsch/Gähler/Straumann 1984, eqn.(5.14), (5.16)). The repercussions of this generousness remains unclear to me (ES).

  30. 30.

    Dirac presented his proposal for a retake of Weyl geometry at the occasion of the symposium honouring his 70th birthday, 1972 at Trieste. This talk remained unpublished. According to (Charap/Tait 1974, p. 249 footnote) the talk was close to his 1973 publication. For the broader historical context of this enterprise, the background in Dirac’s reflection on large numbers in the 1920s, and a surprising link to geophysics see (Kragh 2016).

  31. 31.

    The qualifications “sign inverted” and “generally accepted” refer to the sign convention which agrees with the coordinate-free definition Riem(Y, Z) X = ∇ Y Z X −∇ Z Y X −∇[Y,Z] X. It is preferred in the mathematical literature, including (Weyl 1918b, 5th ed., 131), and also used in the majority of the more recent physics books. The “sign inverted” convention in some of the physics literature goes back to Einstein, e.g. (Einstein 1916, 801), who in turn may have followed Ricci and Levi-Civita. It is found in much of the physics literature from the first half of the 20th century, (Eddington 1923, § 37), (Pauli 1921) up to the influential (Weinberg 1972, eqn. (2.1.3)). Weyl, on the other hand, used the above convention long before the coordinate-free definition of the Riemann tensor was available. It seems to be dominant in the more recent literature on GRT, although Rindler speaks of a 50 % distribution among the two conventions (Rindler 2006, 219).

  32. 32.

    In his discussion with Einstein Weyl had argued that atomic clocks somehow adapt to the local field constellation via the Weylian scalar curvature (letter to Einstein April 28, 1918 etc (Einstein 1998, vol. 8B, doc. 526, 619)), and similarly in the fourth edition of Raum - Zeit - Materie (Weyl 1922, 308f). In the fifth edition Weyl pondered the possibility that the Bohr theory of the atom might give a clue for such an adaptation (Weyl 1918b, 5-th ed., p. 298); compare (Scholz 2017, sec. 5.4).

  33. 33.

    On the work of Canuto and Maeder, see (Kragh 2006, pp. 126ff)

  34. 34.

    For an illuminating historical report on the rise of dark matter, see (Sanders 2010). From a methodological point of view, Maeder’s hypothesis was not so far from the later, more pragmatic and more successful approach of modified Newtonian dynamics, MOND, as introduced by Mordechai Milgrom.

  35. 35.

    Rosen’s “meson” was a hypothetical massive fundamental boson, not a bound state of quarks like the ones in the SM.

  36. 36.

    This was more than a year before the Trieste symposium at which Dirac talked about his ideas. Apparently the paper remained unknown to Dirac. A second paper by Omote followed after Dirac’s publication and after Utiyama had jumped in (Omote 1974).

  37. 37.

    He referred to A. Bregman’s paper, discussed in Section 11.3.3. In Utiyama’s first paper of 1973, Omote went unmentioned; he appears, however, in the references of (Utiyama 1975a).

  38. 38.

    Dirac included a similar scale curvature term in his Lagrangian, but did not study its consequences.

  39. 39.

    Apparently Hayashi/Kugo used different signature conventions from Utiyama, which resulted in another sign flip in ε.

  40. 40.

    In fact, ϕ and φ were not even coupled to the electromagnetic field, as Hayashi and Shirafuri showed in another paper that same year.

  41. 41.

    (Hehl 1970; Trautman 1972, 1973; Hehl et al. 1976b). For ex-post surveys see (Trautman 2006; Hehl 2017) and the rich reader (Blagojević/Hehl 2013).

  42. 42.

    Notations have been slightly adapted.

  43. 43.

    For \(g_{\mu \nu } = h_{\mu }^{\;\;a} h_{\nu a}\) the convention (11.21) boils down to \(g_{\mu \nu } \mapsto g^{\prime }_{\mu \nu } = \varOmega ^2 g_{\mu \nu }\).

  44. 44.

    Bregman, like many other authors, used a sign inverted convention for the scale connection form.

  45. 45.

    Cf. (Sharpe 1997).

  46. 46.

    The torsion term \(\frac {2}{3} T^{\alpha }_{\mu \alpha } \partial ^{\mu }\sigma ^2\) in Bregman’s equation is not scale invariant by itself, but his entire Lagrangian density is scale invariant.

  47. 47.

    This was characteristic for the time. Over several decades the knowledge about the invariance properties of Lagrangian field theories and the know-how of dealing with it spread with marginal or no reference at all to Noether’s seminal paper (Noether 1918), cf. (Kosmann-Schwarzbach 2011).

  48. 48.

    \( \mathfrak {T}^k_{\mu }= \frac {\partial \mathfrak {L}}{\partial h^{\mu }_k}\, , \quad \mathfrak {S}^{\mu }_{ij}= -2 \frac {\partial \mathfrak {L}}{\partial A^{i j}_{\mu }}\, , \quad \mathfrak {D}^{\mu }= \frac {\partial \mathfrak {L}}{\partial A_{\mu }}\) (Charap/Tait 1974, p. 256, notation slightly changed). Compare the dynamical currents by Hehl et al. below (11.29). For the dynamical matter energy current variational and partial derivatives usually coincide. Charap and Tait may have (wrongly) generalized this property to the other currents.

  49. 49.

    (Charap/Tait 1974, eqn. (3.22)).

  50. 50.

    All of them are mentioned in (Blagojević/Hehl 2013, chap. 8).

  51. 51.

    In particular F. Hehl and his various coauthors were active in this field, (Hehl et al. 1976a, 1988a,,b, 1989, 1995). More papers by other authors, though only a few of those just mentioned, appear in (Blagojević/Hehl 2013, chap. 9).

  52. 52.
  53. 53.

    According to later estimates the torsion-spin coupling of Einstein-Cartan gravity becomes important only close to 1038 times the density of a neutron star, which signifies energy densities at the hypothetical grand unification scale of elementary particle interactions (Trautman 2006, p. 194), (Blagojević/Hehl 2013, p. 108). Moreover, there seems to be a problem in defining the spin density of gauge particles, in particular of gluons. This would lead to leaving a large part of the nucleon spin underdetermined (Kleinert 2008, sec. 20.2) (I thank H. Ohanian for this information).

  54. 54.

    (Bohm 1952a,b)

  55. 55.

    Bohm realized the kinship of his approach to the earlier proposals of de Broglie only after he had finished his manuscript (Bohm 1952a, p. 167). In a footnote added in proof he also referred to Madelung’s “similar” approach of 1926, adding the remark “…but like de Broglie he did not carry this interpretation to a logical conclusion” (ibid.).

  56. 56.

    “Die hydrodynamischen Gleichungen sind also gleichwertig mit denen von Schrödinger und liefern alles, was jene geben, d. h. sie sind hinreichend, um die wesentlichen Momente der Quantentheorie der Atome modellmäßig darzustellen” (Madelung 1926, p. 325).

  57. 57.

    (Passon 2015, 2004; Dürr et al. 2009), I thank O. Passon for his helpful explanations of Bohmian mechanics. For relativistic generalizations, see, e.g., (Nicolic 2005). For recent critical remarks see Chen (2016).

  58. 58.

    Cf. (Nicolic 2005, p. 554 ) for vanishing em potential.

  59. 59.

    According to de Broglie it was J.-P. Vigier who made him aware of a parallel between his hypothesis and the work of Einstein and Grommer (de Broglie 1960, p. 92).

  60. 60.

    For E. Nelson’s program to re-derive the quantum dynamics from classical stochastic processes and classical probability see (Bacciagaluppi 2005).

  61. 61.

    See, e.g., (Woodhouse 1991) or (Hall 2001, chaps. 22/23). A classical monograph on the symplectic approach to classical mechanics is (Abraham/Marsden 1978); but this does not discuss spinning particles. In the 1980s the symplectic approach was already used as a starting platform for (pre-)quantization to which proper quantization procedures could then hook up, see e.g. (Śniatycki 1980). Souriau was an early advocate of this program. In his book he discussed relativistic particles with spin (Souriau 1970, §14).

  62. 62.

    Carlos Castro later added his mother’s name Perelman to his second name. Under this name he is mentioned in the acknowledgements of (Santamato 1984b). At that time he was a research assistant at the University of Texas, Austin, where he acquired his Ph.D. in 1991.

  63. 63.

    Santamato and De Martini promoted the geometrical side of their research under different headings: at first they talked about “affine quantum mechanics (AQM)” in “conformal differential geometry” (Santamato/DeMartini 2013), then they shifted to “conformal quantum geometrodynamics (CQG)” (De Martini/Santamato 2014a,b). Physical problems for their approach, like those of negative probability densities for relativistic particles, remain outside the scope of this article.

  64. 64.

    That is, ω ∈ Ω, the sample space of a probability triple \((\varOmega ,\mathfrak {F}, P)\), where \(\mathfrak {F}\subset \mathfrak {P}(\varOmega )\) are the random events and P is a probability measure on Ω.

  65. 65.

    L has the same Euler-Lagrange equations as the original L.

  66. 66.

    Cf. fn 31.

  67. 67.

    Compare Section 11.6.4.1.

  68. 68.

    A sign error in the formula of the Weyl-geometric affine connection (De Martini/Santamato 2014a, eqn. (4)) notwithstanding.

  69. 69.

    (De Martini/Santamato 2014a, p. 3310)

  70. 70.

    For van der Waerdens spinor symbolism see (Schneider 2011).

  71. 71.

    This term, \( \frac {e^2 a^2}{c^2}(H^2-E^2)\), is comparable with the linear term in the field strengths only if the latter are very large, E ∼ 1018 Vm −1, H ∼ 109 T. “To have an idea how large is this field, an electron at rest is accelerated by such field up to 109 GeV  in a linear accelerator 1 m long” (De Martini/Santamato 2014a, p. 3315).

  72. 72.

    It remains unclear to me (E.S.) how the representation matrices of the “Euler angles” of configuration space are suppressed, while the change of coordinate frames in Minkowski space gets represented on the spinor fields.

  73. 73.

    Among others, this had led to the measurement problem for atomic clocks.

  74. 74.

    In the physics literature, and also in the paper by Shojai and Golshani, Ω is often referred to as the “conformal” degree of freedom of the metric, or even the “conformal structure”.The latter is clearly mistaken, whereas the first is, at best, misleading. Therefore I avoid this terminology in favour of scaling degree of freedom.

  75. 75.

    (Shojai/Golshani 1998; Shojai et al. 1998a,b; Shojai/Shojai 2000, 2001).

  76. 76.

    The authors differentiated between “scale transformations” and “conformal transformations”. In their terminology the first operated only on the metric, while the latter rescaled all physical fields according to their weights.

  77. 77.

    The Dirac Lagrangian was stripped of the Yang-Mills term of the scale connection (Shojai/Shojai 2004, eqn. 6).

  78. 78.

    The claim that one can identify a Dirac-type scalar field with a “quantum mass” field remains, in my view, an unfounded speculation; E.S.

  79. 79.

    The authors even conventionalized this idea as a “conformal equivalence principle”(Shojai/Shojai 2003, p. 10), (Shojai/Shojai 2004, p. 63).

  80. 80.

    Not “S. Carroll”, as listed on occasion in the bibliography of later papers.

  81. 81.

    See, e.g. Karaca (2013).

  82. 82.

    The motivation for considering n ≥ 4 was the method of dimensional regularization for the quantization of the theory.

  83. 83.

    In his bibliography he went back directly to (Weyl 1922) and (Weyl 1918a); he did not quote any of the later literature on Weyl geometry.

  84. 84.

    Signs have to be taken with caution. They may depend on conventions for defining the Riemann curvature, the Ricci contraction, and the signature. Smolin, e.g., used a different sign convention for Riem to the one used in this survey. Signs given here are adapted to signature g = (3,  1). The Riemann tensor and Ricci contraction are those usually adopted in the mathematical literature, see fn. 31.

  85. 85.

    This seems to have been widely known. For an explicit statement see, e.g., (Hehl et al. 1996).

  86. 86.

    (Englert et al. 1975) was not quoted by Smolin.

  87. 87.

    In scalar-field gauge with ϕϕ o  = F, his reduced Lagrangian (square gravitational terms dropped) was (Smolin 1979, eqn. (3.17))

  88. 88.

    It is possible to choose the scale gauge such that ϕ becomes constant (scalar-field gauge, see Section 11.2.1).

  89. 89.

    For a survey of the status of investigations in 1981 see (Adler 1982); but note in particular (Zee 1982, 1983). In fact, Zee’s first publication on the subject preceded Smolin’s. (Zee 1979) was submitted in December 1978 and published in February 1979; (Smolin 1979) was submitted in June 1979. The topic of “origin of spontaneous symmetry breaking” by radiative correction was much older (Borrelli 2015; Karaca 2013). A famous paper was (Coleman/Weinberg 1973).

  90. 90.

    Flato/Ra̧cka’s paper appeared as a preprint of the Scuola Internazionale Superiore di Studi Avanzati, Trieste, in 1987; the paper itself was submitted in December 1987 to Physics Letters B and published in July 1988. Cheng’s paper was submitted in February 1988, published in November. Only a decade later, in March 2009, Drechsler and Tann got acquainted with the other two papers. This indicates that the Weyl-geometric approach in field theory had not yet acquired the coherence of a research program with a stable communication network.

  91. 91.

    Personal communication to ES, April 04, 2017.

  92. 92.

    In the sequel the isospin extended scalar field will be denoted by Φ.

  93. 93.

    Cheng added another coupling coefficient for the scale connection, which is here suppressed.

  94. 94.

    See footnote 31.

  95. 95.

    The second term in (11.82) is missing in Cheng’s publication. That is probably not intended, but a misprint. Moreover he did not discuss scale weights for Dirac matrices in the tetrad approach.

  96. 96.

    Remember that the φ terms of scale-covariant derivatives in the Lagrangian of spinor fields cancel.

  97. 97.

    More than a decade earlier Flato had sketched a covariant (“curved space”) generalization of the Wightman axioms (Flato/Simon 1972), different from the one discussed by R. Wald in this volume.

  98. 98.

    (Pawłowski/Ra̧czka 1994a, 1995a,b)

  99. 99.

    For example (Drechsler/Mayer 1977).

  100. 100.

    (Tann 1998, eqn. (372)), (Drechsler 1999, eqn. (2.29)). Both authors used coefficients as in the case of conformal coupling in Riemannian geometry, \(\beta =\frac {1}{6}\). In the Weyl-geometric framework this was an unnecessary restriction but it suppressed the mass factor at the Planck scale for the Weyl field φ. Tann wrote the modified Hilbert term with a negative sign, because he used the sign inverted convention for the Riemann tensor, see footnote 31.

  101. 101.

    (Tann 1998, eqn. (372)), (Drechsler 1999, eqn. (2.46)).

  102. 102.

    In the appendix Drechsler and Tann showed that the squared Weyl-geometric conformal curvature C 2 = C λμνρ C λμνρ arises from the conformal curvature of the Riemannian component by adding a scale curvature term: (Drechsler/Tann 1999, (A 54)). So one may wonder, why they did not replace the square term \( \mathcal {L}_{R^2}\) by the Weyl-geometric conformal curvature term \(\mathcal {L}_{\text{conf}} = \tilde {\alpha } C^2 \sqrt {|det\, g|}\).

  103. 103.

    (11.82) can equivalently be written with a Weylianized scale-covariant derivative \(\overline {D}_{\mu } = \left ( \partial _{\mu } + i \tilde {\varGamma } _{\mu } + w(\psi ) \varphi _{\mu } + \frac {i q}{\hbar c} A_{\mu } \right )\). Because φ μ is real, the scale connection terms w(ψ)φ μ in the Lagrangian cancel.

  104. 104.

    One could then just as well consider a complex valued connection z = (z μ ) with values \(z_{\mu }= \varphi _{\mu } + \frac {i}{\hbar c} A_{\mu }\) in \(\mathbb C= \mathfrak {Lie}(\mathbb C^{\ast })\) and weight \(W(\psi )= (-\frac {3}{2}, q)\). Then \(D_{\mu } \psi = (\partial _{\mu } + \tilde {\varGamma }_{\mu } + W(\psi )z_{\mu })\psi \), presupposing an obvious convention for applying W(ψ)z.

  105. 105.

    This argument is possible, but not compelling: γ|Φ| has the correct scaling weight of mass and may be considered as such.

  106. 106.

    A similar approach is already used in Tann’s PhD dissertation.

  107. 107.

    Note that one could just as well do without (11.84) and proceed with fully scale-covariant masses – compare last footnote.

  108. 108.

    In the literature it is often also called “unitary gauge”.

  109. 109.

    Mathematically speaking, it amounts to a change of trivialization of the SU(2) × U(1)-bundle.

  110. 110.

    \(W^{\pm }_{\mu } = \frac {1}{\sqrt {2}} (W^1_{\mu } \mp i W^2 _{\mu })\), \(Z_{\mu }= \cos \varTheta \, W^3_{\mu } - \sin \varTheta \, B_ {\mu } \).

  111. 111.

    One has to be careful, however. Things become more complicated if one considers the trace. In fact, tr T Φ contains a mass terms of the Dirac field of form \(\gamma | \varPhi _o | \hat {\psi }^{\ast } \hat {\psi }\), with γ coupling constant of the Yukawa term (\(\hat {\psi }\;\) indicating electromagnetic gauge). One of the obstacles for making quantum matter fields compatible with classical gravity is the vanishing of tr T ψ , in contrast to the (nonvanishing) trace of the energy momentum tensor of classical matter. Might Drechsler’s analysis indicate a way out of this impasse? – Warning: The mass-like expressions for W and Z in (11.87) cancel in tr T Φ (Drechsler 1999, eqn. (3.55)) like in the energy-momentum tensor of the W and Z fields themselves.

  112. 112.

    See, e.g., (Franklin 2017).

  113. 113.

    For an exception, still standing under the spell of Drechsler/Tann though with a consistently scale-covariant approach and no explicit scale symmetry breaking term, see (Scholz 2011a).

  114. 114.

    In the high energy physical context, and accordingly in our section 3, signature of g = (+ −−−). For sign conventions regarding curvature see footnote 31.

  115. 115.

    See also (Blagojević 2002, p. 81).

  116. 116.

    For an explicit form of Dirac kinetic terms and Yukawa mass terms see, e.g., (Nishino/Rajpoot 2009, eqn. (1.2)).

  117. 117.

    The expression for the scalar curvature is given in the paper (and also in the later papers by the same authors) as , where a coupling constant f introduced by the authors is here set to f = 1 and transcribed into our notation, (Nishino/Rajpoot 2004, eqn. (14)). The Weyl-geometric value (11.2) in our sign conventions would be , cf. (Weyl 1918c, p. 21), (Drechsler/Tann 1999, eqn. (A 31)) and others. Nishino and Rajpoot apparently used inverted sign conventions for the curvature and the scale connection.

  118. 118.

    The non-broken U(1) symmetry is important for the BRST relations, the quantum analogue of the Noether relations. See (Ruegg et al. 2003, 75ff).

  119. 119.

    The factor \(\zeta _1^{-\frac {1}{2}}\) in (Nishino/Rajpoot 2009, eqn. (2.1)) was set by them to ζ 1 = 1 while transforming the Lagrangian into their eqn. (2.3). The follow up paper (Nishino/Rajpoot 2011, second paragraph of section 2) shows that this reduction was intended. Of course, a different factor ζ 1 would heavily influence the mass calculation in (11.101).

  120. 120.

    Warning: Nishino/Rajpoot used the notation φ for the Stueckelberg “compensator”, i.e. our β, and S μ for the scale connection (the potential of the “Weyl field”), our φ μ . In order to avoid confusion the notation in the present paper has been homogenized for the authors discussed here.

  121. 121.

    Nishino/Rajpoot did not consider the contribution of the modified Hilbert term, in contrast to Smolin and Cheng (see Section 11.5.1).

  122. 122.

    Compare (Stoeltzner 2014).

  123. 123.

    Strictly speaking, their framework does not contain any meaningful “Jordan frame” because their Weyl structure is not integrable, and thus the purely Riemannian representation of the affine connection presupposed for an ordinary Jordan frame does not exist. The Einstein frame, on the other hand, is meaningful in any Weyl-geometric gravity approach with a scale-covariant scalar field and corresponds to scalar-field gauge (11.92).

  124. 124.

    “Note that the Weyl rescaling we made is a field re-definition, but it is not a part of any local scale transformation which is defined to act not only on g μν but also on Φ and φ as in (2.2) [the equation for the full gauge transformation, E.S.]” (Nishino/Rajpoot 2011, p. 4).

  125. 125.

    (Cai/Wei 2007, Acknowledgments)

  126. 126.

    Note that the differential form b = b μ dx μ is sign inverted in comparison with our conventions of Section 11.2.1.

  127. 127.

    Ohanian preserved the label “scale transformation” for a global usage in Minkowski space, where, in addition to the rescaling of the fields \(X\mapsto \tilde {X} = \varOmega ^k X\), a space dilation \(x\mapsto \tilde {x}= \varOmega x\) is applied (Ohanian 2016, p. 25).

  128. 128.

    \(\partial _{\nu }\left (\sqrt {|g|}f^{\mu \nu } \right ) =\mathfrak {J}^{\mu }\). In Ohanian’s Lagrangian φ couples only to the “dilaton” scalar field χ. This leads to a form for the variation of the Lagrangian under scale transformations such that the dynamical current coincides with the Noether current (Ohanian 2016, eqn. (13)).

  129. 129.

    See C.Will’s contribution to this volume.

  130. 130.

    For extensive surveys of this field see (Fujii/Maeda 2003; Capozziello/Faraoni 2011).

  131. 131.

    For a historical discussion of oscillating models see (Kragh 2009) and, in an even wider perspective, H. Kragh’s contribution to this book.

  132. 132.

    The authors declared geodesic incompleteness as “an artifact of an unsuitable frame choice: geodesically incomplete solutions in Einstein frame may be completed in other frames, even though the theories are entirely equivalent away from the singularity” (Bars et al. 2014, p. 13).

  133. 133.

    Isrealit died in 2015 at the age of 87. His last paper known to me is (Israelit 2012), a slightly changed version of (Israelit 2010).

  134. 134.

    Because of scale invariance there is, in fact, only one true scalar field degree of freedom (compare Section 11.2.1).

  135. 135.

    (Israelit 1996, 1999a, 2002a,b); chapters 6 and 7 in his book (Israelit 1999b).

  136. 136.

    Compare on this point (Capozziello/Faraoni 2011, pp. 86ff).

  137. 137.

    J.E. Madriz Aguilar was a student of C. Romero of the Brazilian school, see Section 11.6.4.

  138. 138.

    A few years later high precision numerical modelling showed that thermal effects can completely account for the observations known as the flyby anomaly of the Pioneer spacecrafts (Rievers/Lämmerzahl 2011).

  139. 139.

    A similar argument was already given by Rosen (see Section 11.3.2.2) and more recently by (Romero et al. 2012, sec. 7); compare (Perlick 1989, chap. 5).

  140. 140.

    It was the topic of the author’s talk at the Mainz conference.

  141. 141.

    Bekenstein, as late as 2004, took it as “evident” that measurements with “ clocks and rods” are expressed by the Jordan metric. Moreover, its Levi-Civita connection was taken to govern the free fall of test particles, although the dynamics in the Jordan frame does not satisfy the “usual Einstein equation” (because of explicit terms in the scalar field). The Einstein frame represented for him the “primitive metric” because here the gravitational action reduces to the classical form of the Hilbert term, and the dynamics is given by the Einstein equation (Bekenstein 2004, p. 5f). In their common paper, Milgrom and Bekenstein used the terminology of “dual descriptions” working in “gravitational units” (Einstein frame) respectively “atomic units” (Jordan frame), which sounded a bit like Dirac’s distinction (Bekenstein/Milgrom 1984, p. 14).

  142. 142.

    In his later review paper Bekenstein qualified this point by stating that only so long as the scalar field “…contributes comparatively little to the energy-momentum tensor, it cannot affect light deflection, which will thus be due to the visible matter alone” (Bekenstein 2004, p. 6). One can read this observation the other way round: If the scalar field carries a considerable contribution to the energy-momentum this influences light deflection.

  143. 143.

    Cf. the preceding footnote.

  144. 144.

    The data for 2 clusters are outliers, already from the phenomenological point of view.

  145. 145.

    The famous Coma cluster which led Zwicky to introduce the hypothesis of dark matter is among those for which the Weyl-geometric model gives results consistent with most recent empirical data on mass distributions and accelerations.

  146. 146.

    Unpublished calculations indicate scenarios of a cosmic evolution in agreement with many features of standard cosmology. These would support such features as an initial singularity, large parts of the cosmological redshift due to the expansion of spatial folia in Einstein gauge, accelerated “late time” expansion etc.

  147. 147.

    In this paper Novello still thought in terms of Weyl’s first interpretation of the scale connection, the em dogma in the terminology above.

  148. 148.

    In this paper ω(x) was not yet introduced as a scalar field of its own, but via the square of the electromagnetic potential A μ , i.e., \(\omega =\log {A_{\mu }A^{\mu }}\).

  149. 149.

    In many papers of the Brazilian tradition (Novello et al. 1992) is quoted as a starting point: (Salim/Sautú 1996; de Oliveira 1997; Romero et al. 2011, 2012), to cite just a few. Sometimes it is even called the “first approach to scalar-tensor theory in WIST” [Weyl integrable space-time] Pucheu et al. (2016).

  150. 150.

    In a side remark the authors reminded that − 2ξω λ ω λ is a variationally equivalent kinetic term because the difference to the kinetic term in (11.114) is a total divergence (Novello et al. 1992, p. 654).

  151. 151.

    WIST was (and is) the abbreviation, preferred by the Brazilian authors, for “Weyl integrable spacetime”.

  152. 152.

    This adds flavour to the hybrid approach mentioned above.

  153. 153.

    The fluid Lagrangian was taken from (Ray 1972).

  154. 154.

    For the Hawking-Penrose singularity theorem the geometrical energy condition is the crucial point. It has to be distinguished from the strong energy condition in the physical sense which is given by \(T_{\mu \nu } - \frac {1}{2} tr\, T\, g_{\mu \nu }V^{\mu } V^{\nu }\). The geometrical and physical conditions are equivalent in Einstein gravity; see, e.g., (Curiel 2017, p. 49). Later investigations of the Brazilian school would show that the geometrical energy condition could be violated, while the physical energy condition might still be satisfied (see end of section 5.4.3).

  155. 155.

    (Romero et al. 2011, eq. (7)),(Romero et al. 2012, eqn. (12))

  156. 156.

    “An important conclusion (…) is that general relativity can perfectly ‘survive’ in a non-Riemannian environment” (Romero et al. 2011). In a personal communication with ES (April 3, 2017) C. Romero added “Philosophically, this could perhaps be interpreted as an illustration of the tenets put forward by Poincaré in his conventionalist ideas”.

  157. 157.

    (Pucheu et al. 2014, eqn. (3.24)), (Almeida et al. 2014, eqn. (16)); compare (11.114) in the light of fn. 150.

  158. 158.

    One only needed to put the JBD Lagrangians in a Weyl-geometric framework. Alternatively, if one wants to start from the Brazilian point of view, one may read the constant coefficient of the Hilbert term in (11.119) as the value of a scale-covariant scalar field χ in (Einstein-) scalar-field gauge, χ o ≐ 1, and μ ϕ∂ μ ϕ as the scalar field gauged expression of the scale-covariant kinetic term D μ χD μ χ with scale-covariant derivative (11.3).

  159. 159.

    Only for the authors of (Hehl et al. 1988c) did this appear differently, see Section 11.3.3.

  160. 160.

    Cf. C. Will’s contribution to this volume.

  161. 161.

    E.g. (Drechsler/Hartley 1994).

  162. 162.

    Studies of QFT on Weylian manifolds, comparable to the corresponding researches for Lorentzian manifolds, as discussed in R. Wald’s contribution to this volume, are still a desideratum.

  163. 163.

    These two authors referred neither to the Dirac tradition in Weyl-geometric gravity nor to Utiyama’s; their Weyl-geometric starting point was “self-made” (Drechsler/Hartley 1994) aside from Weyl’s original papers.

  164. 164.

    (Bars et al. 2014, p. 2)

References

  • Abraham, R., & Marsden, J. E. (1978). Foundations of classical mechanics. Redwood City: Addison Wesley. 5-th revised ed. 1985.

    Google Scholar 

  • Adler, S. (1982). Einstein gravity as a symmetry-breaking effect in quantum field theory. Reviews of Modern Physics, 54, 729–766.

    Article  MathSciNet  Google Scholar 

  • Almeida, T. S., Pucheu, M. L., Romero, C., & Formiga, J. B. (2014). From Brans-Dicke gravity to a geometrical scalar-tensor theory. Physical Review D, 89, 064047. arXiv:1311.5459.

    Google Scholar 

  • Audretsch, J. (1983). Riemannian structure of space-time as a consequence of quantum mechanics. Physcial Review D, 27, 2872–2884.

    Article  MathSciNet  Google Scholar 

  • Audretsch, J., Gähler, F., & Straumann, N. (1984). Wave fields in Weyl spaces and conditions for the existence of a preferred pseudo-riemannian structure. Communications in Mathematical Physics, 95, 41–51.

    Google Scholar 

  • Audretsch, J., Hehl, F. W., & Lämmerzahl, C. (1992). Matter wave interferometry and why quantum objects are fundamental for establishing a gravitational theory. In J. Ehlers & G. Schaefer (Eds.), Relativistic gravity research with emphasis on experiments and observations (vol. 410, pp. 369–407). Lecture notes in physics. Berlin: Springer.

    Google Scholar 

  • Audretsch, J., & Lämmerzahl, C. (1988). Constructive axiomatic approach to spacetime torsion. Classical and Quantum Gravity, 5, 1285–1295.

    Article  MathSciNet  MATH  Google Scholar 

  • Audretsch, J., & Lämmerzahl, C. (1991). Establishing the Riemannian structure of space-time by means of light rays and free matter waves. Journal of Mathematical Physics, 32, 2099–2105.

    Article  MathSciNet  MATH  Google Scholar 

  • Audretsch, J., & Lämmerzahl, C. (1994). A new constructive axiomatic scheme for the geometry of space-time. In U. Majer & H.-J. Schmidt (Eds.), Semantical Aspects of Spacetime Theories (pp. 21–40). Mannheim: BI-Verlag.

    Google Scholar 

  • Bacciagaluppi, G. (2005). A conceptual introduction to Nelson’s mechanics. In R. Buccheri, A. Elitzur, & M. Saniga (Eds.), Endophysics, time, quantum and the subjective (pp. 367–388). Singapore: World Scientific. Revised postprint in philsci-archive.pitt.edu/8853/1/Nelson-revised.pdf.

    Google Scholar 

  • Bacciagaluppi, G., & Valentini, A. (2009). Quantum theory at the crossroads. Reconsidering the 1927 solvay conference. Cambridge: Cambridge University Press.

    Google Scholar 

  • Bars, I., Steinhardt, P., & Turok, N. (2014). Local conformal symmetry in physics and cosmology. Physical Review D, 89, 043515. arXiv:1307.1848.

    Google Scholar 

  • Bekenstein, J. (2004). Relativistic gravitation theory for the modified Newtonian dynamics paradigm. Physical Review D, 70, 083509.

    Google Scholar 

  • Bekenstein, J., & Milgrom, M. (1984). Does the missing mass problem signal the breakdown of Newtonian gravity? Astrophysical Journal, 286, 7–14.

    Article  Google Scholar 

  • Bergmann, P. G. (1942). Introduction to the theory of relativity. Englewood Cliffs: Prentice Hall. Reprint New York: Dover 1976.

    Google Scholar 

  • Blagojević, M. (2002). Gravitation and gauge symmetries. Bristol/Philadelphia: Institute of Physics Publishing.

    Google Scholar 

  • Blagojević, M., & Hehl, F. W. (2013). Gauge theories of gravitation. A reader with commentaries. London: Imperial College Press.

    Google Scholar 

  • Bohm, D. (1952a). A suggested interpretation of the quantum theory in terms of ‘hidden variables’ I. Physical Review, 85(1), 166–179.

    Article  MathSciNet  MATH  Google Scholar 

  • Bohm, D. (1952b). A suggested interpretation of the quantum theory in terms of ‘hidden variables’ II. Physical Review, 85(2), 180–193.

    Article  MathSciNet  MATH  Google Scholar 

  • Borrelli, A. (2015). The story of the Higgs boson: the origin of mass in early particle physics. European Physical Journal H, 40(1), 1–52.

    Article  Google Scholar 

  • Bouvier, P., & Maeder, A. (1977). Consistency of Weyl’s geometry as a framework for gravitation. Astrophysics and Space Science, 54, 497–508.

    Article  MathSciNet  MATH  Google Scholar 

  • Brans, C. (1961). Mach’s principle and a varying gravitational constant. PhD thesis, Physics Department, Princeton University

    Google Scholar 

  • Brans, C. (1999). Gravity and the tenacious scalar field. In A. Harvey (Ed.), On Einstein’s path. Essays in honor of engelbert schücking (pp. 121–138). Berlin: Springer. arXiv:gr-qc/9705069.

    Chapter  Google Scholar 

  • Brans, C. (2005). The roots of scalar-tensor theories: an approximate history. In International workshop on gravitation and cosmology, Contributions to the Cuba Workshop, Santa Clara 2004. arXiv:gr-qc/0506063.

    Google Scholar 

  • Brans, C. (2014). Jordan-Brans-Dicke theory. Scholarpedia, 9(4), 31358.

    Article  Google Scholar 

  • Brans, C., & Dicke, R. H. (1961). Mach’s principle and a relativistic theory of gravitation. Physical Review, 124, 925–935.

    Article  MathSciNet  MATH  Google Scholar 

  • Bregman, A. (1973). Weyl transformations and Poincaré gauge invariance. Progress of Theoretical Physics, 49, 667–6992.

    Article  Google Scholar 

  • Cai, R.-G., & Wei, H. (2007). Cheng-Weyl vector field and its cosmological application. Journal of Cosmology and Astroparticle Physics, 0709, 015. arXiv:astro-ph/0607064.

    Google Scholar 

  • Calderbank, D., & Pedersen, H. (1998). Einstein-Weyl geometry. Advances in Mathematics, 97, 74–109.

    Google Scholar 

  • Callan, C., Coleman, S., & Jackiw, R. (1970). A new improved energy-momentum tensor. Annals of Physics, 59, 42–73.

    Article  MathSciNet  MATH  Google Scholar 

  • Canuto, V., Adams, P. J., Hsieh S.-H., & Tsiang, E. (1977). Scale covariant theory of gravitation and astrophysical application. Physical Review D, 16, 1643–1663.

    Article  MathSciNet  MATH  Google Scholar 

  • Canuto, V., & Goldman, I. (1983). Astrophyiscal consequences of a violation of the strong equivalence principle. Nature, 304, 311–314.

    Article  Google Scholar 

  • Capozziello, S., & Faraoni, V. (2011). Beyond Einstein gravity. A survey of gravitational theories for cosmology and astrophysics. Dordrecht: Springer.

    Google Scholar 

  • Carroll, R. (2004). Gravity and the quantum potential. arXiv:gr-qc/0406004.

    Google Scholar 

  • Castro, C. (1992). On Weyl geometry, random processes, and geometric quantum mechanics. Foundations of Physics, 22, 569–615.

    Article  MathSciNet  Google Scholar 

  • Castro, C. (2007). On dark energy, Weyl’s geometry, different derivations of the vacuum energy density and the Pioneer anomaly. Foundations of Physics, 37, 366–409.

    Article  MathSciNet  MATH  Google Scholar 

  • Castro, C. (2009). The cosmological constant and Pioneer anomaly from Weyl geometry and Mach’s principle. Physics Letters B, 675, 226–230.

    Article  MathSciNet  Google Scholar 

  • Charap, J. M., & Tait, W. (1974). A gauge theory of the Weyl group. Proceedings Royal Society London A, 340, 249–262.

    Article  MathSciNet  MATH  Google Scholar 

  • Chen, P., & Kleinert, H. (2016). Deficiencies of Bohm trajectories in view of basic quantum principles. Electronic Journal of Theoretical Physics, 13(35), 1–12.

    Google Scholar 

  • Cheng, H. (1988). Possible existence of Weyl’s vector meson. Physical Review Letters, 61, 2182–2184.

    Article  Google Scholar 

  • Clifton, T., Ferreira, P., Padilla, A., & Skordis, C. (2012). Modified gravity and cosmology. Physics Reports, 513, 1–189. arXiv:1106.2476.

    Article  MathSciNet  Google Scholar 

  • Codello, A., D’Orodico, G., Pagani, C., & Percacci, R. (2013). The renormalization group and Weyl invariance. Classical and Quantum Gravity, 30, 115015. arXiv:1210.3284.

    Article  MathSciNet  MATH  Google Scholar 

  • Coleman, R., & Korté, H. (1984). Constraints on the nature of inertial motion arising from the universality of free fall and the conformal causal structure of spacetime. Journal of Mathematical Physics, 25, 3513–3526.

    Article  MathSciNet  MATH  Google Scholar 

  • Coleman, S., & Weinberg, E. (1973). Radiative corrections as the origin of sponteneous symmetry breaking. Physical Review D, 7, 1888–1910.

    Article  Google Scholar 

  • Cotsakis, S., & Miritzis, J. (1999). Variational and conformal structure of nonlinear metric-connection gravitational Lagrangians. Journal of Mathematical Phyiscs, 40(6), 3063.

    Article  MathSciNet  MATH  Google Scholar 

  • Curiel, E. (2017). A primer on energy conditions. In (Lehmkuhl et al. 2017, pp. 43–104).

    Google Scholar 

  • Dahia, F., Gomez, A. T., & Romero, C. (2008). On the embedding of space-time in five-dimensional Weyl spaces. Journal of Mathematical Physics, 49, 102501. arXiv:0711.2754.

    Article  MathSciNet  MATH  Google Scholar 

  • Darrigol, O. (2014). Physics and necessity: Rationalist pursuits from the cartesian past to the quantum present. Oxford: Oxford University Press.

    Google Scholar 

  • de Broglie, L. (1960). Non-linear wave mechanics. A causal interpretation. Amsterdam: Elsevier. Translated by A.J. Knodel and J.C. Miller.

    Google Scholar 

  • De Martini, F., & Santamato, E. (2014a). Interpretation of quantum-nonlocality by conformal geometrodynamics. International Journal of Theoretical Physics, 53, 3308–3322. arXiv:1203:0033.

    Google Scholar 

  • De Martini, F., & Santamato, E. (2014b). Nonlocality, no-signalling, and Bell’s theorem investigated by Weyl conformal differential geometry. Physica Scripta, 2014, T163. arXiv:1406.2970.

    Article  Google Scholar 

  • De Martini, F., & Santamato, E. (2014c). The intrinsic helicity of elementary particles and the spin-statistic connection. International Journal of Quantum Information, 12, 1560004.

    Article  MathSciNet  MATH  Google Scholar 

  • De Martini, F., & Santamato, E. (2015). Proof of the spin-statistics theorem. Foundations of Physics, 45(7), 858–873.

    Article  MathSciNet  MATH  Google Scholar 

  • De Martini, F., & Santamato, E. (2016). Proof of the spin-statistics theorem in the relativistic regime by Weyl’s conformal quantum mechanics. International Journal of Quantum Information, 14(04), 1640011.

    Article  MathSciNet  MATH  Google Scholar 

  • de Oliveira, H. P., Salim, J. M., & Sautú, S. L. (1997). Non-singular inflationary cosmologies in Weyl integrable spacetime. Classical and Quantum Gravity, 14(10), 2833–2843.

    Article  MathSciNet  MATH  Google Scholar 

  • Deser, S. (1970). Scale invariance and gravitational coupling. Annals of Physics, 59, 248–253.

    Article  MathSciNet  Google Scholar 

  • Dicke, R. H. (1962). Mach’s principle and invariance under transformations of units. Physical Review, 125, 2163–2167.

    Article  MathSciNet  MATH  Google Scholar 

  • Dirac, P. A. M. (1973). Long range forces and broken symmetries. Proceedings Royal Society London A, 333, 403–418.

    Article  MathSciNet  Google Scholar 

  • Dirac, P. A. M. (1974). Cosmological models and the large number hypothesis. Proceedings Royal Society London A, 338, 439–446.

    Article  Google Scholar 

  • Drechsler, W. (1999). Mass generation by Weyl symmetry breaking. Foundations of Physics, 29, 1327–1369.

    Article  MathSciNet  Google Scholar 

  • Drechsler, W., & Hartley, D. (1994). The role of the internal metric in generalized Kaluza-Klein theories. Journal of Mathematical Physics, 35, 3571–3585.

    Article  MathSciNet  MATH  Google Scholar 

  • Drechsler, W., & Mayer, M. E. (1977). Fibre bundle techniques in gauge theories. Lectures in mathematical physics at the University of Austin (vol. 67) Lecture notes in physics. Berlin: Springer.

    Google Scholar 

  • Drechsler, W., & Tann, H. (1999). Broken Weyl invariance and the origin of mass. Foundations of Physics, 29(7), 1023–1064. arXiv:gr-qc/98020.

    Article  MathSciNet  Google Scholar 

  • Dürr, D., Goldstein, S., Tumulka, R., & Zanghi N. (2009). Bohmian mechanics. In D. Greenberger, K. Hentschel, & F. Weinert (Eds.), Compendium of quantum physics (pp. 47–55). Berlin: Springer.

    Chapter  Google Scholar 

  • Eddington, A. S. (1923). The mathematical theory of relativity. Cambridge: Cambridge University Press.

    Google Scholar 

  • Ehlers, J., Pirani, F., & Schild, A. (1972). The geometry of free fall and light propagation. In L. O’Raifertaigh (Ed.), General relativity, papers in honour of J.L. Synge (pp. 63–84). Oxford: Clarendon Press.

    Google Scholar 

  • Einstein, A. (1916). Die Grundlagen der allgemeinen Relativitätstheorie. Mathematische Annalen, 49, 769–822.

    Article  MATH  Google Scholar 

  • Einstein, A. (1949). Autobiographical notes (vol. 7) The library of living philosophers. La Salle, IL: Open Court.

    Google Scholar 

  • Einstein, A. (1998). The collected papers of Albert Einstein. Volume 8. The Berlin years: Correspondence, 1914–1918, Part B: 1918. R. Schulmann, A. J. Kox, M. Janssen, J. Illy, & K. von Meyenn (Eds.). Princeton: Princeton University Press.

    Google Scholar 

  • Englert, F., Gunzig, E., Truffin, C., & Windey, P. (1975). Conformal invariant relativity with dynamical symmetry breakdown. Physics Letters, 57 B, 73–76.

    Google Scholar 

  • Englert, F., & Truffin, C. (1976). Conformal invariance in quantum gravity. Nuclear Physics B, 117, 407–432.

    Article  Google Scholar 

  • Faraoni, V., & Nadeau, S. (2007). (Pseudo)issue of the conformal frame revisited. Physical Review D, 75(2), 023501.

    Google Scholar 

  • Flato, M., & Raçka, R. (1988). A possible gravitational origin of the Higgs field in the standard model. Physics Letters B, 208, 110–114. Preprint, SISSA (Scuola Internazionale Superiore di Studi Avanzate), Trieste, 1987 107/87/EP.

    Article  Google Scholar 

  • Flato, M., & Simon, J. (1972). Wightman formulation for the quantization of the gravitational field. Physical Review D, 5, 332–341.

    Article  MathSciNet  Google Scholar 

  • Folland, G. B. (1970). Weyl manifolds. Journal of Differential Geometry, 4, 145–153.

    Article  MathSciNet  MATH  Google Scholar 

  • Foot, R., & Kobakhidze, A. (2013). Electroweak scale invariant models with small cosmological constant. International Journal of Modern Physics A, 30(21), 1550126. arXiv:0709.2750.

    Article  MathSciNet  MATH  Google Scholar 

  • Foot, R., Kobakhidze, A., McDonald, K., & Volkas, R. R. (2007a). Neutrino mass in radiatively-broken scale-invariant models. Physical Review D, 76, 075014. arXiv:0706.1829.

    Google Scholar 

  • Foot, R., Kobakhidze, A., McDonald, K., & Volkas, R. R. (2007b). A solution to the hierarchy problem from an almost decoupled hidden sector within a classically scale invariant theory. Physical Review D, 77, 035006. arXiv:0709.2750.

    Google Scholar 

  • Franklin, A. (2017). The missing piece of the puzzle: The discovery of the Higgs boson. Synthese, 194(2), 259–274. https://doi.org/10.1007/s11229-014-0550-y.

    Article  MathSciNet  Google Scholar 

  • Fujii, Y., & Maeda, K.-C. (2003). The scalar-tensor theory of gravitation. Cambridge: Cambridge University Press.

    Google Scholar 

  • Gauduchon, P. (1995). La 1-forme de torsion d’une variété hermitienne compacte. Journal für die reine und angewandte Mathematik, 469, 1–50.

    Google Scholar 

  • Gilkey, P., Nikcevic, S., & Simon, U. (2011). Geometric realizations, curvature decompositions, and Weyl manifolds. Journal of Geometry and Physics, 61, 270–275. arXiv:1002.5027.

    Article  MathSciNet  MATH  Google Scholar 

  • Goenner, H. (2004). On the history of unified field theories. Living Reviews in Relativity, 7, 2. http://relativity.livingreviews.org/Articles/lrr-2004-2.

  • Goenner, H. (2012). Some remarks on the genesis of scalar-tensor theories. General Relativity and Gravity, 44(8), 2077–2097. arXiv: 1204.3455.

    Article  MathSciNet  MATH  Google Scholar 

  • Gray, J. (Ed.) (1999). The symbolic universe: Geometry and physics 1890–1930. Oxford: Oxford University Press.

    Google Scholar 

  • Hall, B. C. (2001). Quantum theory for mathematicians. Berlin: Springer.

    Google Scholar 

  • Hayashi, K., & Kugo, T. (1979). Remarks on Weyl’s gauge field. Progress of Theoretical Physics, 61, 334–346.

    Article  Google Scholar 

  • Hehl, F. W. (1970). Spin und Torsion in der allgemeinen Relativitätstheorie oder die Riemann-Cartansche Geometrie der Welt. Habilitationsschrift. Technische Universität Clausthal: Mimeograph.

    Google Scholar 

  • Hehl, F. W. (2017). Gauge theory of gravity and spacetime. In (Lehmkuhl et al. 2017, 145–170).

    Chapter  Google Scholar 

  • Hehl, F. W., Kerlick, G. D., & von der Heyde, P. (1976a). On a new metric affine theory of gravitation. Physcis Letters B, 63 (4), 443–448.

    Article  MathSciNet  Google Scholar 

  • Hehl, F. W., McCrea, J. D., & Kopczyǹski, W. (1988a). The Weyl group and ist currents. Physics Letters A, 128, 313–318.

    Article  MathSciNet  Google Scholar 

  • Hehl, F. W., McCrea, J. D., & Mielke, E. (1988b). Skaleninvarianz und Raumzeit-Struktur. In B. Geyer, H. Herwig, & H. Rechenberg (Eds.), Werner Heisenberg. Physiker und Philosoph (pp. 299–306). Berlin: Spektrum.

    Google Scholar 

  • Hehl, F. W., McCrea, J. D., Mielke, E., & Ne’eman, Y. (1989). Progress in metric-affine theories of gravity with local scale invariance. Foundations of Physics, 19, 1075–1100.

    Article  MathSciNet  Google Scholar 

  • Hehl, F. W., McCrea, J. D., Mielke, E., & Ne’eman, Y. (1995). Metric-affine gauge theory of gravity: Field equations, noether identities, world spinors, and breaking of dilation invariance. Physics Reports, 258, 1–171.

    Article  MathSciNet  Google Scholar 

  • Hehl, F. W., Mielke, E., & Tresguerres, R. (1988c). Weyl spacetimes, the dilation current, and creation of gravitating mass by symmetry breaking. In W. Deppert & K. Hübner (Eds.), Exact sciences and their philosophical foundations; exakte wissenschaften und ihre philosophische grundlegung (pp. 241–310). Frankfurt: Peter Lang.

    Google Scholar 

  • Hehl, F. W., Puntigam, R., & Tsantilis, E. (1996). A quadratic curvature Lagrangian of Pawlowski and Raczka: A finger exercise with MathTensor. In F. W. Hehl, R. Puntigam, & H. Ruder (Eds.), Relativity and scientific computing…. Berlin: Springer. [gr-qc/9601002].

    Google Scholar 

  • Hehl, F. W., von der Heyde, P., Kerlick, G. D., & Nester, J. M. (1976b). General relativity with spin and torsion: Foundations and prospects. Reviews of Modern Physics, 48, 393–416.

    Article  MathSciNet  MATH  Google Scholar 

  • Higa, T. (1993). Weyl manifolds and Einstein-Weyl manifolds. Commentarii Mathematici Sancti Pauli, 42, 143–160.

    Google Scholar 

  • Israelit, M. (1996). Conformally coupled dark matter. Astrophysics and Space Science, 240(1), 331–344. arXiv:gr-qc/9608035.

    Article  MathSciNet  MATH  Google Scholar 

  • Israelit, M. (1999a). Matter creation by geometry in an integrable Weyl– Dirac theory. Foundations of Physics, 29(8), 1303–1322.

    Google Scholar 

  • Israelit, M. (1999b). The Weyl-Dirac theory and our universe. New York: Nova Science.

    Google Scholar 

  • Israelit, M. (2002a). Primary matter creation in a Weyl-Dirac cosmological model. Foundation of Physics, 32, 295–321.

    Google Scholar 

  • Israelit, M. (2002b). Quintessence and dark matter created by Weyl-Dirac geometry. Foundation of Physics, 32, 945–961.

    Google Scholar 

  • Israelit, M. (2010). A Weyl-Dirac cosmological model with DM and DE. General Relativity and Gravitationi, 43, 751–775. arXiv:1008.0767.

    Article  MathSciNet  MATH  Google Scholar 

  • Israelit, M. (2012). Nowadays cosmology with the Weyl-Dirac approach. Preprint arXiv:1212.2208. Slightly changed version of Israelit (2010).

    Google Scholar 

  • Israelit, M., & Rosen, N. (1992). Weyl-Dirac geometry and dark matter. Foundations of Physics, 22, 555–568.

    Article  MathSciNet  Google Scholar 

  • Israelit, M., & Rosen, N. (1993). Weylian dark matter and cosmology. Foundations of Physics, 24, 901–915.

    Article  Google Scholar 

  • Israelit, M., & Rosen, N. (1995). Cosmic dark matter and Dirac gauge function. Foundations of Physics, 25, 763–777.

    Article  Google Scholar 

  • Jordan, P. (1952). Schwerkraft und weltall. Braunschweig: Vieweg. 2nd revised edtion 1955.

    Google Scholar 

  • Kaiser, D. (2006). Whose mass is it anyway? Particle cosmology and the objects of a theory. Social Studies of Science, 36(4), 533–564.

    Article  MathSciNet  Google Scholar 

  • Kaiser, D. (2007). When fields collide. Scientific American, pp. 62–69.

    Google Scholar 

  • Karaca, K. (2013). The construction of the Higgs mechanism and the emergence of the electroweak theory. Studies in History and Philosophy of Modern Physics, 44, 1–16.

    Article  MathSciNet  MATH  Google Scholar 

  • Kasuya, M. (1975). On the gauge theory in the Einstein-Cartan-Weyl space-time. Nuovo Cimento B, 28(1), 127–137.

    Article  MathSciNet  Google Scholar 

  • Kibble, T. (1961). Lorentz invariance and the gravitational field. Journal for Mathematical Physics, 2, 212–221. In (Blagojević/Hehl 2013, chap. 4).

    Article  MathSciNet  MATH  Google Scholar 

  • Kleinert, H. (2008). Multivalued fields in condensed matter, electromagnetism, and gravitation. Singapore: World Scientific.

    Google Scholar 

  • Kosmann-Schwarzbach, Y. (2011). The noether theorems. invariance and conservation laws in the twentieth century. Berlin: Springer.

    Google Scholar 

  • Kostant, B. (1970). Quantization and unitary representations. 1. Prequantisation (vol. 170) Lecture notes in mathematics. Berlin: Springer.

    Google Scholar 

  • Kragh, H. (1999). Quantum generations: A history of physics in the twentieth century. Princeton: Princeton University Press.

    Google Scholar 

  • Kragh, H. (2006). Cosmologies with varying speed of light: A historical perspective. Studies in History and Philosophy of Modern Physics, 37, 726–737.

    Article  MathSciNet  MATH  Google Scholar 

  • Kragh, H. (2009). Continual fascination: The oscillating universe in modern cosmology. Science in Context, 22(4), 587–612.

    Article  Google Scholar 

  • Kragh, H. (2016). Varying gravity. Dirac’s legacy in cosmology and geophyics. Science networks. Heidelberg: Springer-Birkhäuser.

    Google Scholar 

  • Lämmerzahl, C. (1990). The geometry of matter fields. In V. de Sabbata & J. Audretsch (Eds.), Quantum mechanics in curved spacetime (pp. 23–48). Berlin: Springer.

    Google Scholar 

  • Lehmkuhl, D. (2014). Why Einstein did not believe that general relativity geometrizes gravity. Studies in History and Philosophy of Modern Physics, 46B, 316–326. http://philsci-archive.pitt.edu/9825/.

    Article  MathSciNet  MATH  Google Scholar 

  • Lehmkuhl, D., Schiemann, G., & Scholz, E. (Eds.) (2017). Towards a theory of spacetime theories. Einstein studies. Berlin/Basel: Springer/Birkhäuser.

    Google Scholar 

  • Lobo, I. I., Barreto, A. B., & Romero, C. (2015). Space-time singularities in Weyl manifolds. European Physics Journal C, 75 (9), 448. arXiv:1506.02180.

    Google Scholar 

  • Madelung, E. (1926). Quantentheorie in hydrodynamischer Form. Zeitschrift für Physik, 40(3–4), 322–326.

    Google Scholar 

  • Maeder, A. (1978a). Metrical connection in space-time, Newton’s and Hubble’s law. Astronomy and Astrophysics, 65, 337–343.

    Google Scholar 

  • Maeder, A. (1978b). Cosmology II: Metrical connection and clusters of galaxies. Astronomy and Astrophysics, 67, 81–86.

    Google Scholar 

  • Mannheim, P. (1994). Open questions in classical gravity. Foundations of Physics, 224, 487–511.

    Article  MathSciNet  Google Scholar 

  • Mannheim, P. (2000). Attractive and repulsive gravity. Foundations of Physics, 22, 709–746.

    Article  MathSciNet  Google Scholar 

  • Mannheim, P. (2012). Making the case for conformal gravity. Foundations of Physics, 42. arXiv:1101.2186.

    Article  MathSciNet  MATH  Google Scholar 

  • Mannheim, P., & Kazanas, D. (1989). Exact vacuum solution to conformal Weyl gravity and galactic rotation curves. Astrophysical Journal, 342, 635–638.

    Article  MathSciNet  Google Scholar 

  • Meissner, K., & Nicolai, H. (2009). Conformal symmetry and the standard model. Physics Letters B, 648, 312–317. arXiv:hep-th/0612165.

    Article  MathSciNet  MATH  Google Scholar 

  • Miritzis, J. (2004). Isotropic cosmologies in Weyl geometry. Classical and Quantum Gravity, 21, 3043–3056. arXiv:gr-qc/0402039.

    Article  MathSciNet  MATH  Google Scholar 

  • Miritzis, J. (2013a). Energy exchange in Weyl geometry. In Proceedings of the Greek Relativity Meeting NEB15, June 2012, Chania, Greece. Journal of physics: Conference series. arXiv:1301.5402.

    Google Scholar 

  • Miritzis, J. (2013b). Acceleration in Weyl integrable spacetime. International Journal of Modern Physics D, 22(5), 1350019. arXiv:1301.5696.

    Article  Google Scholar 

  • Myvold, W. (2003). On some early objections to Bohm’s theory. International Studies in the Philsophy of Science, 17(1), 7–24.

    Article  MathSciNet  MATH  Google Scholar 

  • Narlikar, J., & Padmanabhan, T. (1983). Quantum cosmology via path integrals. Physics Reports, 100, 151–200.

    Article  MathSciNet  Google Scholar 

  • Nicolic, H. (2005). Relativistic quantum mechanics and the Bohmian interpretation. Foundations of Physics Letters, 18(6), 549–561.

    Article  MathSciNet  MATH  Google Scholar 

  • Nieh, H.-T. (1982). A spontaneously broken conformal gauge theory of gravitation. Physics Letters A, 88, 388–390.

    Article  Google Scholar 

  • Nishino, H., & Rajpoot, S. (2004). Broken scale invariance in the standard model. arXiv:hep-th/0403039.

    Google Scholar 

  • Nishino, H., & Rajpoot, S. (2007). Broken scale invariance in the standard mode. AIP Conference Proceedings, 881, 82–93. arXiv:0805.0613 (with different title).

    Google Scholar 

  • Nishino, H., & Rajpoot, S. (2009). Implication of compensator field and local scale invariance in the standard model. Physical Review D, 79, 125025. arXiv:0906.4778.

    Google Scholar 

  • Nishino, H., & Rajpoot, S. (2011). Weyl’s scale invariance for standard model, renormalizability and zero cosmological constant. Classical and Quantum Gravity, 28, 145014.

    Article  MathSciNet  MATH  Google Scholar 

  • Noether, E. (1918). Invariante variationsprobleme. Göttinger nachrichten pp. 235–257. In Gesammelte abhandlungen (vol. 1, pp. 770ff). Berlin: Springer.

    Google Scholar 

  • Novello, M. (1969). Dirac’s equation in a Weyl space. Nuovo Cimento A, 94(4), 954–960.

    Article  MathSciNet  Google Scholar 

  • Novello, M., & Heintzmann, H. (1983). Weyl integrable space-time: A model for the cosmos? Physics Letters A, 98(1), 10–11.

    Article  Google Scholar 

  • Novello, M., Oliveira, L. A. R., Salim, J. M., & Elbaz, E. (1992). Geometrized instantons and the creation of the universe. International Journal of Modern Physics D, 1, 641–677.

    Article  MathSciNet  MATH  Google Scholar 

  • Obukhov, Y. (1982). Conformal invariance and space-time torsion. Physics Letters A, 90, 13–16.

    Article  MathSciNet  Google Scholar 

  • Ohanian, H. (2016). Weyl gauge-vector and complex dilaton scalar for conformal symmetry and its breaking. General Relativity and Gravity, 48(25). https://doi.org/10.1007/s10714-016-2023-8. arXiv:1502.00020.

  • Omote, M. (1971). Scale transformations of the second kind and the Weyl space-time. Lettere al Nuovo Cimento, 2(2), 58–60.

    Article  MathSciNet  Google Scholar 

  • Omote, M. (1974). Remarks on the local-scale-invariant gravitational theory. Lettere al Nuovo Cimento, 10(2), 33–37.

    Article  MathSciNet  Google Scholar 

  • O’Raifeartaigh, L. (1997). The dawning of gauge theory. Princeton: Princeton University Press.

    Google Scholar 

  • Ornea, L. (2001). Weyl structures in quaternionic geomety. A state of the art. In E. Barletta (Ed)., Selected topics in geometry and mathematical physics (vol. 1, pp. 43–80). Potenza: University of degli Studi della Basilicata. arXiv:math/0105041.

    Google Scholar 

  • Padmanabhan, T. (1989). Quantum cosmology – the story so far. In B. R. Iyer, N. Mukunda, & C. V. Vishveshwara (Eds.), Gravitation, gauge theories and the early universe (pp. 373–404). Dordrecht: Kluwer

    Chapter  Google Scholar 

  • Passon, O. (2004). Bohmsche Mechanik. Eine Einführung in die determinisitsche Interpretation der Quantenmechanik. Frankfurt/Main: Harri Deutsch.

    Google Scholar 

  • Passon, O. (2015). Nicht-Kollaps-Interpretationen der Quantentheorie. In C. Friebe, M. Kuhlmann, H. Lyre, P. M. Näger, O. Passon, & M. Stöckler (Eds.), Philosophie der quantenphysik (pp. 178–224). Berlin: Springer.

    Google Scholar 

  • Pauli, W. (1921). Relativitätstheorie. In Encyklopädie der Mathematischen Wissenschaften (vol. 5, pp. 539–775). Leipzig: Teubner. Collected Papers I, 1–237. Reprint edited and commented by D. Giulini, Berlin etc. Springer 2000.

    Google Scholar 

  • Pauli, W. (1940). Über die Invarianz der Dirac’schen Wellengleichungen gegenüber Ähnlichkeitstransformationen des Linienelementes im Fall verschwindender Ruhmasse. Helvetia Physica Acta, 13, 204–208. In (Pauli 1964, II, 918–922).

    Google Scholar 

  • Pauli, W. (1964). Collected scientific papers, R. Kronig, V. F. Weisskopf (Eds.). New York: Wiley.

    Google Scholar 

  • Pawłowski, M. (1990). Can gravity do what the Higgs does? Preprint IC/90/454.

    Google Scholar 

  • Pawłowski, M., & Ra̧czka, R. (1994a). Mass generation in the standard model without dynamical Higgs field. Preprint. hep-th/9403303.

    Google Scholar 

  • Pawłowski, M., & Ra̧czka, R. (1994b). A unified conformal model for fundamental interactions without dynamical Higgs field. Foundations of Physics, 24, 1305–1327. ILAS 4/94 hep-th/9407137.

    Article  Google Scholar 

  • Pawłowski, M., & Ra̧czka, R. (1995a). A Higgs-free model for fundamental interactions and its implications. Preprint. ILAS/EP-1-1995.

    Chapter  Google Scholar 

  • Pawłowski, M., & Ra̧czka, R. (1995b). A Higgs-free model for fundamental interactions. Part I: Formulation of the model. In J. Bertrand, M. Flato, J.-P. Gazeau, M. Irac-Astaud, & D. Sternheimer (Eds.), Modern group theoretical methods in physics (pp. 221–232). Springer Science+Business Media: Dordrecht. Preprint ILAS/EP-3-1995, hep-ph/9503269.

    Google Scholar 

  • Penrose, R. (1965). Zero rest-mass fields including gravitation: asymptotic behaviour. Proceedings Royal Society London A, 284, 159–203.

    Article  MathSciNet  MATH  Google Scholar 

  • Penrose, R. (2006). Before the big bang: An outrageous perspective and ist implications for partivle physics. In Proceedings of EPAC 2006, Edinburgh, Scotland. http://accelconf.web.cern.ch/accelconf/e06/PAPERS/THESPA01.PDF.

  • Perlick, V. (1987). Characterization of standard clocks by means of light rays and freely falling particles’. General Relativity and Gravitation, 19, 1059–1073.

    Article  MathSciNet  MATH  Google Scholar 

  • Perlick, V. (1989). Zur Kinematik Weylscher Raum-Zeit-Modelle. Dissertationsschrift TU Berlin.

    Google Scholar 

  • Perlick, V. (1991). Observer fields in Weylian spacetime models. Classical and Quantum Gravity, 8, 1369–1385.

    Article  MathSciNet  MATH  Google Scholar 

  • Pfister, H. (2004). Newton’s first law revisited. Foundations of Physics Letters, 17, 49–64.

    Article  MathSciNet  MATH  Google Scholar 

  • Pickering, A. (1988). Constructing quarks. Edinburgh: Edinburgh University Press.

    Google Scholar 

  • Pucheu, M. L., Almeida, T. S., & Romero, C. (2014). A geometrical approach to Brans-Dicke theory. In C. M. Gonzales, J. E. M. Aguliar, & L. M. R. Barrera (Eds.), Accelerated Cosmic Expansion. Astrophysics and space science proceedings (vol. 38, pp. 33–41). Berlin: Springer

    Google Scholar 

  • Pucheu, M. L., Alves, F. A. P., Barreto, A. B., & Romero, C. (2016). Cosmological models in Weyl geometric scalar tensor theory. Physical Review D, 6, 064010. arXiv:1602.06966.

    Google Scholar 

  • Quiros, I. (2000a). Dual geometries and spacetime singularities. Physical Review D, 61, 124026.

    Google Scholar 

  • Quiros, I. (2000b). Transformations of units and world’s geometry. Preprint. gr-qc/0004014.

    Google Scholar 

  • Quiros, I. (2013). Scale invariance and broken electroweak symmetry may coexist together. Preprint. arXiv:1312.1018.

    Google Scholar 

  • Quiros, I. (2014a). Scale invariance: fake appearances. Preprint. arXiv:1405.6668.

    Google Scholar 

  • Quiros, I. (2014b). Scale invariant theory of gravity and the standard model of particles. Preprint. arXiv:1401.2643.

    Google Scholar 

  • Quiros, I., Bonal, R., & Cardenas, R. (2000). Brans-Dicke-type theories and avoidance of the cosmological singularity. Physical Review D, 62, 044042.

    Google Scholar 

  • Quiros, I., Garcìa-Salcedo, R., Madriz Aguilar, J., & Matos, T. (2013). The conformal transformations’ controversy: what are we missing. General Relativity and Gravitation, 45, 489–518. arXiv:1108.5857.

    Google Scholar 

  • Ray, J. (1972). Lagrangian density for perfect fluids in general relativity. Journal of Mathematical Physics, 13(10), 1451–1453.

    Article  MATH  Google Scholar 

  • Rievers, B., & Lämmerzahl, C. (2011). High precision thermal modeling of complex systems with application to the flyby and Pioneer anomaly. Annalen der Physik, 532(6), 439. arXiv:1104.3985.

    Article  MATH  Google Scholar 

  • Rindler, W. (2006). Relativity. Special, General, and Cosmological. Oxford: Oxford University Press. 2nd ed. 2007.

    Google Scholar 

  • Romero, C., Fonseca-Neto, J.B., & Pucheu, M. L. (2011). General relativity and Weyl frames. International Journal of Modern Physics A, 26(22), 3721–3729. arXiv:1106.5543.

    Article  MathSciNet  MATH  Google Scholar 

  • Romero, C., Fonseca-Neto, J. B., & Pucheu, M. L. (2012). General relativity and Weyl geometry. Classical and Quantum Gravity, 29 (15), 155015. arXiv:1201.1469.

    Article  MathSciNet  MATH  Google Scholar 

  • Rosen, N. (1982). Weyl’s geometry and physics. Foundations of Physics, 12, 213–248.

    Article  MathSciNet  Google Scholar 

  • Ruegg, H., & Ruiz-Altaba, M. (2003). The Stueckelberg field. Preprint. arXiv:hep-th/0304245.

    Google Scholar 

  • Ryckman, T. (2005). The reign of relativity. Philosophy in physics 1915–1925. Oxford: Oxford University Press.

    Google Scholar 

  • Salim, J. M., & Sautú, S. L. (1996). Gravitational theory in Weyl integrable spacetime. Classical and Quantum Gravity, 13 (2), 363–360.

    Article  MathSciNet  MATH  Google Scholar 

  • Sanders, R. H. (2010). The dark matter problem. A historical perspective. Cambridge: Cambridge University Press.

    Google Scholar 

  • Santamato, E. (1984a). Geometric derivation of the Schrödinger equation from classical mechanics in curved Weyl spaces. Physical Review D, 29, 216–222.

    Article  MathSciNet  Google Scholar 

  • Santamato, E. (1984b). Statistical interpretation of the Klein-Gordon equation in terms of the spacetime Weyl curvature. Journal of Mathematical Physics 25(8), 2477–2480.

    Article  MathSciNet  Google Scholar 

  • Santamato, E. (1985). Gauge-invariant statistical mechanics and average action principle for the Klein-Gordon particle in geometric quantum mechanics. Physical Review D, 32(10), 2615–2621.

    Article  MathSciNet  Google Scholar 

  • Santamato, E., & De Martini, F. (2013). Derivation of the Dirac equation by conformal differential geometry. Foundations of Physics, 43(5), 631–641. arxiv:1107.3168.

    Article  MathSciNet  MATH  Google Scholar 

  • Schneider, M. (2011). Zwischen zwei Disziplinen. B.L. van der Waerden und die Entwicklung der Quantenmechanik. Berlin: Springer.

    Chapter  Google Scholar 

  • Scholz, E. (1999). Weyl and the theory of connections. In Gray (1999). pp. 260–284.

    Google Scholar 

  • Scholz, E. (Ed.) (2001). Hermann Weyl’s Raum - Zeit - Materie and a general introduction to his scientific work. Basel: Birkhäuser.

    Google Scholar 

  • Scholz, E. (2005a). Einstein-Weyl models of cosmology. In J. Renn (Ed.), Albert Einstein. 100 authors for Einstein (pp. 394–397). Weinheim: Wiley-VCH.

    Google Scholar 

  • Scholz, E. (2005b). On the geometry of cosmological model building. Preprint. arXiv:gr-qc/0511113.

    Google Scholar 

  • Scholz, E. (2009). Cosmological spacetimes balanced by a Weyl geometric scale covariant scalar field. Foundations of Physics, 39, 45–72. arXiv:0805.2557.

    Article  MathSciNet  MATH  Google Scholar 

  • Scholz, E. (2011a). Weyl geometric gravity and electroweak symmetry ‘breaking’. Annalen der Physik, 523, 507–530. arxiv.org/abs/1102.3478.

    Article  MathSciNet  MATH  Google Scholar 

  • Scholz, E. (2011b). Weyl’s scale gauge geometry in late 20th century physics. Preprint. arXiv:1111.3220.

    Google Scholar 

  • Scholz, E. (2016a). MOND-like acceleration in integrable Weyl geometric gravity. Foundations of Physics, 46, 176–208. arXiv:1412.0430.

    Article  MathSciNet  MATH  Google Scholar 

  • Scholz, E. (2016b). Clusters of galaxies in a Weyl geometric approach to gravity. Journal of Gravity, 46, 9706704. arXiv:1506.09138.

    Google Scholar 

  • Scholz, E. (2017). Paving the way for transitions – a case for Weyl geometry. In (Lehmkuhl et al. 2017, pp. 171–224). arXiv:1206.1559.

    Google Scholar 

  • Schouten, J. A. (1924). Der Ricci-Kalkl̈. Eine Einführung in die neueren Methoden und problem der mehrdimensionalen differentialgeometrie. Die grundlehren der mathematischen wisenschaften (vol. 10). Berlin: Springer.

    Google Scholar 

  • Schouten, J. A. (1954). Ricci calculus (2nd ed.). Berlin: Springer.

    Book  MATH  Google Scholar 

  • Sciama, D. W. (1962). On the analogy between charge and spin in general relativity. In Recent developments in general relativity Festschrift for L. Infeld (pp. 415–439). Oxford and Warsaw: Pergamon and PWN. In (Blagojević/Hehl 2013, chap. 4).

    Google Scholar 

  • Shaposhnikov, M., & Zenhäusern, D. (2009a). Quantum scale invariance, cosmological constant and hierarchy problem. Physics Letters B, 671, 162–166. arXiv:0809.3406.

    Article  Google Scholar 

  • Shaposhnikov, M., & Zenhäusern, D. (2009b). “Scale invariance, unimodular gravity and dark energy. Physics Letters B, 671, 187–192. arXiv:0809.3395.

    Article  MathSciNet  Google Scholar 

  • Sharpe, R. W. (1997). Differential geometry: Cartan’s generalization of Klein’s Erlangen program. Berlin: Springer.

    Google Scholar 

  • Shojai, A. (2000). Quantum gravity and cosmology. International Journal of Modern Physics A, 15(2), 1757–1771.

    Article  MathSciNet  MATH  Google Scholar 

  • Shojai, F., & Golshani, M. (1998). On the geometrization of Bohmian quantum mechanics: A new approach to quantum gravity. International Journal of Modern Physics A, 13(4), 677–693.

    Article  MathSciNet  MATH  Google Scholar 

  • Shojai, A., & Shojai, F. (2000). Nonminimal scalar-tensor theories and quantum gravity. International Journal of Modern Physics A, 15(13), 1859–1868.

    Article  MathSciNet  MATH  Google Scholar 

  • Shojai, F., & Shojai, A. (2001). About some problems raised by the relativistic form of de-Broglie-Bohm theory of pilot wave. Physica Scripta, 54(5), 413–416. arXiv:gr-qc/0404102.

    Article  MathSciNet  MATH  Google Scholar 

  • Shojai, F., & Shojai, A. (2003). On the relation of Weyl geometry and Bohmian quantum mechanics. Gravitation and Cosmology, 9(3), 163ff. Max Planck Institute for Gravitational Physics, Preprint AEI-2002-060. arXiv:gr-qc/0306099.

    Google Scholar 

  • Shojai, F., & Shojai, A. (2004). Understanding quantum theory in terms of geometry. Preprint. arXiv:gr-qc/0404102.

    Google Scholar 

  • Shojai, F., Shojai, A., & Golshani, M. (1998a). Conformal transformations and quantum gravity. Modern Physics Letters A, 13 (34), 2725–2729.

    Article  MathSciNet  MATH  Google Scholar 

  • Shojai, F., Shojai, A., & Golshani, M. (1998b). Scalar tensor theories and quantum gravity. Modern Physics Letters A, 13(36), 1915–2922.

    Article  MathSciNet  MATH  Google Scholar 

  • Shojai, A., Shojai, F., & Golshani, M. (1998c). Nonlocal effects in quantum gravity. Modern Physics Letters A, 13(37), 2965–2969.

    Article  MathSciNet  MATH  Google Scholar 

  • Simms, D. (1978). On the Schrödinger equation given by geometric quantization. In K. Bleuler, H. R. Petry, & A. Reetz (Eds.), Differential geometrical methods in mathematical physics II (vol. 676, pp. 351–356) Lecture notes in mathematics. Berlin: Springer.

    Google Scholar 

  • Smolin, L. (1979). Towards a theory of spacetime structure at very short distances. Nuclear Physics B, 160, 253–268.

    Article  MathSciNet  Google Scholar 

  • Śniatycki, J. (1980). Geometric quantization and quantum mechanics. Berlin: Springer.

    Book  MATH  Google Scholar 

  • Souriau, J.-M. (1966). Quantification géométrique. Communications in Mathematical Physics, 1, 374–398.

    Google Scholar 

  • Souriau, J.-M. (1970). Structure des systèmes dynamiques. Paris: Duno. English as (Souriau 1997).

    Google Scholar 

  • Souriau, J.-M. (1997). Structure of dynamical systems. A symplectic view ofh physics. Berlin: Springer. Translated from (Souriau 1970) by C.-H. Cushman-de Vpassonries.

    Chapter  Google Scholar 

  • Steinhardt, P., & Turok, N. (2002). Cosmic evolution in a cyclic universe. Physical Review D, 65(12), 126003. arXiv:hep-th/0111098.

    Google Scholar 

  • Stoeltzner, M. (2014). Higgs models and other stories about mass generation. Journal for the General Philosophy of Science, 45, 369–386.

    Google Scholar 

  • Tann, H. (1998). Einbettung der quantentheorie eines skalarfeldes in eine Weyl geometrie — Weyl symmetrie und ihre brechung. München: Utz.

    Google Scholar 

  • Tonnelat, M.-A. (1965). Les Théories unitaires de l’électromagnétisme et de la gravitation. Paris: Gauthier-Villars.

    MATH  Google Scholar 

  • Trautman, A. (1972). On the Einstein-Cartan equations I, II. Bulletin Academie Polonaise des Sciences, Série des sciences math., astr. et phys., 20, 185–191, 503.

    MathSciNet  Google Scholar 

  • Trautman, A. (1973). On the structure of the Einstein-Cartan equations. Symposia Mathematica, 12, 139–162. Relativitá convegno del Febbraio del 1972.

    Google Scholar 

  • Trautman, A. (2006). Einstein-Cartan theory. In J.-P. Françoise, G. L. Naber, & S. T. Tsou (Eds.), Encyclopedia of mathematical physics (vol. 2, pp. 189–195). Oxford: Elsevier. In (Blagojević/Hehl 2013, chap. 4).

    Chapter  Google Scholar 

  • Trautman, A. (2012). Editorial note to J. Ehlers, F. A. E. Pirani and A. Schild, The geometry of free fall and light propagation. General Relativty and Gravity, 441, 1581–1586.

    Article  MATH  Google Scholar 

  • Utiyama, R. (1956). Invariant theoretical interpretation of interaction. Physical Review, 101(5), 1597–1607.

    Article  MathSciNet  MATH  Google Scholar 

  • Utiyama, R. (1973). On Weyl’s gauge field. Progress of Theoretical Physics, 50, 2028–2090.

    Article  MathSciNet  MATH  Google Scholar 

  • Utiyama, R. (1975a). On Weyl’s gauge field. General Relativity and Gravitation, 6, 41–47.

    Article  MathSciNet  MATH  Google Scholar 

  • Utiyama, R. (1975b). On Weyl’s gauge field II. Progress of Theoretical Physics, 53, 565–574.

    Article  MathSciNet  MATH  Google Scholar 

  • Vizgin, V. (1994). Unified Field Theories in the First Third of the 20th Century. Basel: Birkhäuser. Translated from the Russian by J. B. Barbour.

    Book  MATH  Google Scholar 

  • Weinberg, S. (1972). Gravitation and cosmology. New York: Wiley.

    Google Scholar 

  • Weyl, H. (1918a). Gravitation und Elektrizität. Sitzungsberichte der königlich preußischen akademie der wissenschaften zu Berlin (pp. 465–480). In (Weyl 1968, II, 29–42), English in (O’Raifeartaigh 1997, 24–37).

    Google Scholar 

  • Weyl, H. (1918b). Raum, - Zeit - Materie. vorlesungen über allgemeine relativitätstheorie. Berlin: Springer. Other editions: 21919, 31919, 41921, 51923, 61970, 71988, 81993. English and French translations from the 4th ed. in 1922.

    Google Scholar 

  • Weyl, H. (1918c). Reine Infinitesimalgeometrie. Mathematische Zeitschrift, 2, 384–411. In (Weyl 1968, II, 1–28).

    Article  MathSciNet  MATH  Google Scholar 

  • Weyl, H. (1920). Letter H. Weyl to F. Klein, December 28, 1920. Nachlass F. Klein Universitätsbibliothek Göttingen Codex Ms Klein 12, 297.

    Google Scholar 

  • Weyl, H. (1921). Zur Infinitesimalgeometrie: Einordnung der projektiven und der konformen Auffassung. Nachrichten Göttinger gesellschaft der wissenschaften (pp. 99–112). In (Weyl 1968, II, 195–207).

    Google Scholar 

  • Weyl, H. (1922). Space – Time – Matter.Translated from the 4th German edition by H. Brose. London: Methuen. Reprint New York: Dover 1952.

    Google Scholar 

  • Weyl, H. (1949). Philosophy of Mathematics and Natural Science. Princeton: Princeton University Press. 21950, 32009.

    Google Scholar 

  • Weyl, H. (1949/2016). Similarity and congruence: a chapter in the epistemology of science. ETH Bibliothek, Hs 91a:31. Published in (Weyl 1955, 3rd edition, 153–166).

    Google Scholar 

  • Weyl, H. (1955). Symmetrie. Ins Deutsche übersetzt von Lulu Bechtolsheim. Basel/Berlin: Birkhäuser/Springer. 21981, 3rd edition Ergänzt durch einen Text aus dem Nachlass ‘Symmetry and congruence’, ed. D. Giulini et al. 2016: Springer.

    Google Scholar 

  • Weyl, H. (1968). Gesammelte Abhandlungen.K. Chandrasekharan (Ed.), vol. 4. Berlin: Springer.

    Google Scholar 

  • Wood, W. R., & Papini, G. (1992). Breaking Weyl invariance in the interior of a bubble. Physical Review D, 45, 3617–3627.

    Article  MathSciNet  Google Scholar 

  • Wood, W. R., & Papini, G. (1997). A geometric approach to the quantum mechanics of de Broglie and Vigier. In S. Jeffers, S. Roy, J.-P. Vigier, & G. Hunter (Eds.), The Present Status of the Quantum Theory of Light. Proceedings in the Honour of Jean-Pierre Vigier (pp. 247–258). Dordrecht: Kluwer. arXiv:gr-qc/9612042.

    Chapter  Google Scholar 

  • Woodhouse, N. M. J. (1991). Geometric quantization. Oxford: Clarendon.

    MATH  Google Scholar 

  • Wu, C.-L. (2004). Conformal scaling gauge symmetry and inflationary universe. International Jorunal of Modern Physics A, 20, 811ff. arXiv:astro-ph/0607064.

    Article  MathSciNet  MATH  Google Scholar 

  • Yang, C.-N. (1980). Einstein’s impact on theoretical physics. Physics Today, 33(6), 42–49. In (Yang 1983, 563–567).

    Article  MathSciNet  Google Scholar 

  • Yang, C. N. (1983). Selected papers 1945–1980. With commentary. San Francisco: Freeman.

    MATH  Google Scholar 

  • Yang, C. N., & Mills, R. (1954). Conservation of isotopic spin and isotopic gauge invariance. Physical Review, 96, 191–195. In (Yang 1983, 172–176)

    Article  MathSciNet  MATH  Google Scholar 

  • Yuan, F.-F., & Huang, Y.-C. (2013). A modified variational principle for gravity in Weyl geometry. Classical and Quantum Gravity, 30(19), 195008. arXiv:1301.1316.

    Article  MathSciNet  MATH  Google Scholar 

  • Zee, A. (1979). Broken-symmetric theory of gravity. Physical Review Letters, 42, 417–421.

    Article  Google Scholar 

  • Zee, A. (1982). A theory of gravity based on the Weyl-Eddington action. Physics Letters B, 109, 183–186.

    Article  Google Scholar 

  • Zee, A. (1983). Einstein gravity emerging from quantum Weyl gravity. Annals of Physics, 151, 431–443.

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This paper owes its existence to David Rowe’s initiative in several respects. He encouraged me to present heterodox ideas on Weyl-geometric methods in cosmology at the conference and invited me to rethink the case after a cool reception of the talk by the other participants. That gave me the chance to place my views in the wider range of recent attempts to use Weyl-geometric methods in physics. After an interruption of several years, an earlier first draft of this paper (Scholz 2011b) had be to be rewritten completely for the final version of this book. The new version overlaps nicely with the wider ambit of the investigations of the interdisciplinary group Epistemology of the LHC with center at Wuppertal and supported generously by the DFG/FWF. This group offers the chance for a close communication between historians and philosophers of science and collegues from the elementary particle community. H. Cheng, F. Hehl, J. Miritzis, C. Romero, D. Rowe, A. Trautman, S. Walter gave helpful hints for the final version of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Erhard Scholz .

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer Science+Business Media, LLC, part of Springer Nature

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Scholz, E. (2018). The Unexpected Resurgence of Weyl Geometry in late 20th-Century Physics. In: Rowe, D., Sauer, T., Walter, S. (eds) Beyond Einstein. Einstein Studies, vol 14. Birkhäuser, New York, NY. https://doi.org/10.1007/978-1-4939-7708-6_11

Download citation

  • DOI: https://doi.org/10.1007/978-1-4939-7708-6_11

  • Published:

  • Publisher Name: Birkhäuser, New York, NY

  • Print ISBN: 978-1-4939-7706-2

  • Online ISBN: 978-1-4939-7708-6

  • eBook Packages: Physics and AstronomyPhysics and Astronomy (R0)

Publish with us

Policies and ethics