This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\UseRawInputEncoding

Structure of conformal gravity in the presence of a scale breaking scalar field

Philip D. Mannheim Department of Physics, University of Connecticut, Storrs, CT 06269, USA
[email protected]
(January 17 2022)
Abstract

We revisit the structure of conformal gravity in the presence of a c-number, conformally coupled, long range, macroscopic scalar field. And in the static, spherically symmetric case discuss two classes of exact exterior solutions. In one solution the scalar field has a constant value and in the other solution, which is due to Brihaye and Verbin, it has a dependence on the radial coordinate, with the two exterior solutions being relatable by a conformal transformation. In light of these two solutions Horne and then Hobson and Lasenby raised the concern that the fitting of conformal gravity to galactic rotation curves had been misapplied and thus called the successful fitting of the conformal theory into question. In this paper we show that the analysis of Brihaye and Verbin needs to reappraised, with their reported result not being as general as they had indicated, but nonetheless being valid in the particular case that they studied. For the analyses of Horne and of Hobson and Lasenby we show that this macroscopic scalar field is not related to the mass generation that is required in a conformal theory. Rather, not just in conformal gravity, but also in standard Einstein gravity, the presence of such a long range scalar field would lead to test particles whose masses would be of the same order as the masses of the galaxies around which they orbit. Since particle masses are not at all of this form, such macroscopic fields cannot be responsible for mass generation; and the existence of any such mass-generating scalar fields can be excluded, consistent with there actually being no known massless scalar particles in nature. Instead, mass generation has to be due to c-number vacuum expectation values of q-number fields. Such expectation values are microscopic not macroscopic and only vary within particle interiors, giving particles an extended, baglike structure, as needed for localization in a conformal theory. And being purely internal they have no effect on galactic orbits, to thus leave the good conformal gravity fitting to galactic rotation curves intact.

I Introduction

I.1 Implications of mass generation for gravity

Even though massive sources play a central role in theories of gravity, as long as the gravity theory is both classical and macroscopic, these masses can be treated as being purely mechanical or kinematical in nature. However, the actual generation of mass itself is intrinsically microscopic, being due to a quantum-field-theoretic, spontaneously broken Hilbert space vacuum in which a quantum field such as ϕ\phi or a quantum fermion bilinear such as ψ¯ψ\bar{\psi}\psi acquires a non-vanishing vacuum expectation value. Nonetheless, it is generally thought that macroscopic gravity is not sensitive to such microscopic effects, with gravity only needing to respond to the presence of a massive source irrespective of how that mass may or may not have arisen. And yet once one has an expectation value such as S(x)=Ω|ϕ|ΩS(x)=\langle\Omega|\phi|\Omega\rangle or S(x)=Ω|ψ¯ψ|ΩS(x)=\langle\Omega|\bar{\psi}\psi|\Omega\rangle, one can ask whether this c-number S(x)S(x) order parameter might itself extend to large distances and have macroscopic consequences. And if the vacuum is translation invariant the quantity S(x)S(x) would be a constant and thus would indeed extend throughout all spacetime, and not only would it do so, it would then provide a potentially substantial contribution to the cosmological constant. Moreover, if evaluated in a coherent state |C|C\rangle, a quantity such as S(x)=C|ϕ|CS(x)=\langle C|\phi|C\rangle or S(x)=C|ψ¯ψ|CS(x)=\langle C|\bar{\psi}\psi|C\rangle would then be spacetime dependent and could thus have a long range spacetime dependence. With such a spacetime-dependent S(x)S(x) carrying both energy and momentum it could contribute to the classical gravitational equations of motion in a possibly significant way. In addition, if Yukawa coupled to a fermion via a Lagrangian density term of the form hS(x)ψ¯(x)ψ(x)-hS(x)\bar{\psi}(x)\psi(x), such an S(x)S(x) could modify the motion of a fermionic particle over large distances scales, so that its trajectory (or that of a star composed of fermions) in a gravitational field would then not be geodesic.

To see the nature of the issue in detail consider a minimally coupled scalar field with wave equation μμS=0\nabla_{\mu}\nabla^{\mu}S=0 propagating in a static, spherically symmetric geometry with line element ds2=B(r)dt2A(r)dr2r2dθ2r2sin2θdϕ2ds^{2}=B(r)dt^{2}-A(r)dr^{2}-r^{2}d\theta^{2}-r^{2}\sin^{2}\theta d\phi^{2} in an arbitrary metric theory of gravity. In such a geometry the wave equation would take the form (ABr2)1/2r[r2(B/A)1/2rS]=0(ABr^{2})^{-1/2}\partial_{r}[r^{2}(B/A)^{1/2}\partial_{r}S]=0, and the scalar field would acquire a radial dependence of the form S(r)=c𝑑rA1/2/r2B1/2S(r)=c\int drA^{1/2}/r^{2}B^{1/2} where cc is an integration constant. If the metric coefficients B(r)B(r) and A(r)A(r) are fixed by some static source such as a galaxy with NN^{*} stars so that A(r)A(r) and B(r)B(r) are of order NN^{*}, S(r)S(r) would accordingly acquire a term of order NN^{*}. But that would lead to a Yukawa coupling to a single star orbiting the galaxy that would also be of order NN^{*}. Such a coupling could not be ascribed to the mass generation mechanism for a single star as it would be NN^{*} times too large, and not only that, it would give an effect that would vary from one galaxy to the next as NN^{*} itself would then vary.

Such a situation could for instance have been encountered in the scalar-tensor Brans-Dicke theory Brans1961 , in which a macroscopic scalar field is coupled to Einstein gravity. However, Brans and Dicke only allowed the scalar field to couple to the gravitational sector and not to the matter sector. And as emphasized in Weinberg1972 , any such coupling to the matter sector could impair the successes of the equivalence principle. The Brans-Dicke theory thus finesses the issue by not associating the macroscopic scalar field with mass generation, i.e., the Brans-Dicke scalar field does not Yukawa couple to propagating fermions. In the Brans-Dicke theory the lack of any such Yukawa coupling is essentially imposed by fiat. However, the typical Lagrangian density for a scalar field, viz. [μSμS/2m2S2/2+λS4]-[\nabla_{\mu}S\nabla^{\mu}S/2-m^{2}S^{2}/2+\lambda S^{4}], has a discrete symmetry under SSS\rightarrow-S, and thus imposing this symmetry could actually forbid the presence of any hSψ¯ψ-hS\bar{\psi}\psi term as it would be odd under the discrete symmetry if ψ¯ψ\bar{\psi}\psi is even. Alternatively, the scalar field could be associated with some very high mass scale such as a cosmological scale or a grandunified scale, scales that are not associated with low scale fermion masses.

In not Yukawa coupling the scalar field to the matter sector the Brans-Dicke theory avoids another problem. Specifically, while the c-number S(x)S(x) might describe the vacuum or a coherent state, quantizing fluctuations around either such configuration would give rise to propagating quantum scalar fields, which if Yukawa coupled to fermions would mediate interactions between fermions, in direct analog to photon or graviton exchange. These interactions would then be long range if the quantized scalar field is massless. There appears to be no sign of any such long range interaction, and concomitantly, there actually appears to be no known scalar (or pseudoscalar) particle that is in fact massless. Thus for the SU(2)×U(1)SU(2)\times U(1) electroweak theory for instance, while the vacuum expectation value of the Higgs field gives rise to fermion masses, the associated fluctuating scalar Higgs field that is excited out of the Higgs vacuum is actually massive, to thus lead to acceptably short range rather than long range interactions. In addition, the electroweak Higgs field is accompanied by a set of massless Goldstone bosons, and initially the exchange of these particles could give rise to long range interactions. However, via the Englert-Brout-Higgs mechanism these Goldstone bosons are incorporated into a matching set of gauge bosons that then become massive, with no massless Goldstone bosons being left in the spectrum. The electroweak theory thus nicely protects itself from the presence of any long range interaction that could be mediated by spin zero scalar or pseudoscalar particles, with this all being done at the microscopic, not macroscopic, level.

While these remarks would apply to any metric theory of gravity, for the conformal gravity theory of interest to us in this paper there are additional considerations. Specifically, in a theory that has a conformal symmetry there are no mass scales at the level of the Lagrangian. Consequently, all mass must be generated in the vacuum, and thus both the stars in a galaxy and a test particle that is orbiting around the galaxy must all acquire their masses by one and the same mechanism. Thus unlike in the Brans-Dicke case, one cannot decouple the mass generation mechanism for the stars in a gravitational source (the stars that fix the above A(r)A(r) and B(r)B(r) coefficients) from the mass generation mechanism for the test star that orbits around the galaxy. Thus if the scalar field does couple to the stars in the source, it must couple to the test particle too, and one is not free to ignore this coupling. Moreover, if one were to introduce a discrete SSS\rightarrow-S symmetry, while it might then prevent the scalar field from coupling to the test particle, it would equally prevent the scalar field from coupling to and giving mass to the particles in the source as well.

An additional, and quite unusual aspect of the conformal gravity theory is that even if there is a radially dependent macroscopic scalar field, according to Brihaye2009 it can actually hide itself by not contributing to the energy-momentum tensor in the region exterior to a static, spherically symmetric source. Its presence thus does not modify the exterior region solution for the metric that would have been obtained had it not been present at all. However, such a scalar field is not completely without consequence since, as noted in Horne2016 and Hobson2021 , its Yukawa coupling to a test particle (as realized via a test particle action of the form IT=hS(x)𝑑sI_{T}=-h\int S(x)ds) would lead to a trajectory for the test particle that is not geodesic. Such non-geodesic behavior would represent a challenge to the dark-matter-free conformal gravity fits to 138 galactic rotation curves given in Mannheim2011 ; Mannheim2012 ; O'Brien2012 , fits that were based on the use of geodesic orbits. Thus in contrast to the Brans-Dicke study where the scalar field couples to the geometry but not to the test particle, in the conformal case the scalar field couples to the test particle but does not affect the geometry in the exterior region.

It is the purpose of this paper to revisit the analyses of Brihaye2009 , Horne2016 and Hobson2021 , and show that they actually reinforce the point we raised above, namely that one cannot associate a radially varying, macroscopic scalar field with mass generation in the first place. Now initially this might not appear to be the case since the very fact that the scalar field causes the exterior region TμνT_{\mu\nu} to vanish means that the strength of the scalar field is not fixed by the number NN^{*} of stars in the source. However, it turns out that we need to reexamine the study made in Brihaye2009 . Specifically, for a general static, spherically symmetric energy-momentum tensor there are only three independent components: T00T_{00}, TrrT_{rr} and TθθT_{\theta\theta}. However, for the traceless TμνT_{\mu\nu} required of a conformal theory there are only two, and as will become apparent below, the specific structure of the conformal gravitational equations actually make T00TrrT^{0}_{\phantom{0}0}-T^{r}_{\phantom{r}r} and TrrT^{r}_{\phantom{r}r} the optimal choice. And as noted in Brihaye2009 , the vanishing of T00TrrT^{0}_{\phantom{0}0}-T^{r}_{\phantom{r}r} in the region exterior to the matter source would force a classical, macroscopic scalar field to have the form S(r)=1/(b+ar)S(r)=1/(b+ar); with the very vanishing of T00TrrT^{0}_{\phantom{0}0}-T^{r}_{\phantom{r}r} allowing the parameters aa and bb to not be fixed by any matter fields that are located in the interior region.

Now in their study Brihaye and Verbin and then subsequently Horne and then Hobson and Lasenby took TrrT^{r}_{\phantom{r}r} to vanish as well. However, we shall show below that this form for S(r)S(r) does not in and of itself actually cause TrrT^{r}_{\phantom{r}r} to vanish. That this could in principle be the case is due to the fact that for the general line element ds2=B(r)dt2A(r)dr2r2dθ2r2sin2θdϕ2ds^{2}=B(r)dt^{2}-A(r)dr^{2}-r^{2}d\theta^{2}-r^{2}\sin^{2}\theta d\phi^{2} the covariant conservation of a traceless TμνT_{\mu\nu} leads to

(B2B1r)(T00Trr)(ddr+4r)Trr=0.\displaystyle\left(\frac{B^{\prime}}{2B}-\frac{1}{r}\right)(T^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r})-\left(\frac{d}{dr}+\frac{4}{r}\right)T^{r}_{{\phantom{r}}r}=0. (1.1)

Consequently, at best the vanishing of T00TrrT^{0}_{\phantom{0}0}-T^{r}_{\phantom{r}r} could only constrain TrrT^{r}_{\phantom{r}r} to be of the form A/r4A/r^{4} where AA is a constant. In our study below we will use the full set of equations of motion of the theory to show that TrrT^{r}_{\phantom{r}r} does not in fact vanish if b=0b=0, but does vanish if b0b\neq 0. With b0b\neq 0 we now can set all the components of TμνT_{\mu\nu} to zero and thus establish the result that had been assumed in Brihaye2009 ; Horne2016 ; Hobson2021 where the coefficient bb had been taken to be non-zero. In order to actually get TrrT^{r}_{\phantom{r}r} to vanish we have to use the equations of motion, and this now does lead us to a relation between aa and bb and the metric coefficients produced by the interior region matter source, to thus force aa and bb to be of order NN^{*} in the galactic case. (As noted in Horne2016 , one can obtain a relation between the aa and bb parameters and the metric coefficients through the scalar field equation of motion, but as we show below, this does not force TrrT^{r}_{\phantom{r}r} to vanish.)

While mass generation is microscopic, nonetheless it is of interest to examine what would happen in the conformal theory if mass generation were to lead to a scalar field S(x)S(x) that did vary macroscopically and couple to a test particle with an action of the form IT=hS(x)𝑑sI_{T}=-h\int S(x)ds. In this paper we shall examine this option in detail in the explicitly solvable model introduced in Brihaye2009 , and explicitly demonstrate that such a test particle action would give rise to masses for test stars that were not of order one solar mass but of order the mass of the galaxy around which they orbit. Since it is thus not valid to ascribe mass generation to such a macroscopic field, the concerns raised in Horne2016 and Hobson2021 (studies that used this ITI_{T} test particle action) are thus invalidated.

If mass generation is not to be associated with an S(x)S(x) that varies macroscopically then it must be associated with an S(x)S(x) that only varies on microscopic scales while being constant on large scales. Structures in which the order parameter S(x)S(x) only varies within the particles that it gives mass to are well established in flat spacetime dynamical theories where they are known as extended models or bag models of hadrons, i.e., models of hadrons with non-zero radii. In particular, they can arise in flat spacetime theories that have no intrinsic dimensional parameters. And it was this particular aspect of mass generation that first led Mannheim1994 ; Mannheim1994a the present author to generalize these flat space concepts to curved space in a gravity theory, conformal gravity, in which there again would be no intrinsic dimensional parameters. Moreover, such extended structures are actually necessary in the conformal theory since the mechanism that generates mass dynamically generates extended structures at the same time, to thereby enable particles to localize into the stars and galaxies that are of relevance to astrophysics and cosmology. Moreover, in these dynamical models coherent states such as |C|C\rangle are in and of themselves not actually stable. (They would be stable if they had a non-trivial topology, though that is not the case here.) The coherent states only become stable in the presence of a positive energy fermion excited out of the filled negative energy Dirac sea. And in this stabilization mechanism the fermion becomes localized and the scalar field only varies from a constant value within the localization region. Thus the very nature of dynamical mass generation is such that there can be no macroscopic order parameters that vary radially on large distance scales. This is completely consistent with our showing in this paper that one could not ascribe mass generation to macroscopic scalar fields that do vary on large distance scales.

This paper is organized as follows. After introducing the conformal theory in the rest of this section, in Sec. II we present the exact solution to the conformal theory in the static, spherically symmetric case in which there are no matter fields in the exterior region. In this analysis we present the linear potentials that are key to the conformal gravity theory fitting to galactic rotation curves, noting that there are actually two linear potentials of relevance, one due to the stars inside a local inhomogeneous matter source such as a galaxy and the other being due to the homogeneous global background cosmology. In Sec. III we follow Brihaye2009 and augment the theory with a scalar field that is to be present in this exterior region, and investigate its implications for the behavior of the scalar field TμνT_{\mu\nu}. This analysis enables us to establish that the linear potential remains in the solution and that the aa and bb parameters in the S(r)=1/(b+ar)S(r)=1/(b+ar) solution are of galactic and even cosmological scale. In Sec. IV we discuss the implications of coupling the same scalar field to a test particle. As noted in Horne2016 the linear potential contribution is removed from the test particle orbit in the S=1/(b+ar)S=1/(b+ar) case, while it is not present at all in the S=1/bS=1/b case, with the effect of the linear potential thus being lost in both of the cases. However, in either case we establish that at the same time this would lead to a test particle that would acquire an unacceptable galactic or cosmological sized mass. In addition, we identify some inconsistencies in the calculation given in Horne2016 . Moreover, while, as noted in Horne2016 ; Hobson2021 , one can actually remove the key linear potential by a conformal transformation when there is just a single source and radial symmetry, in Sec. V we show that this is not the case for a multiparticle source such as a spiral galaxy with axial not radial symmetry. Finally, in Sec. VI we show that through use of the microscopic scalar field C|ψ¯ψ|C\langle C|\bar{\psi}\psi|C\rangle that is appropriate to dynamical symmetry breaking test particles then move on standard geodesics. That this is the case is because in the self-consistent vacuum the conformal symmetry is broken in both the vacuum and the action, with the fermion that is excited out of the vacuum not being constrained to propagate according to a conformal invariant equation of motion, and with the linear potential term then not being removed even though the scalar field is macroscopically constant. In an appendix we follow Mannheim2006 and discuss the cosmological origin of the global linear potential, with there being a conformal and coordinate equivalence between the static, spherically symmetric geometries of relevance to galactic rotation curves and the comoving Robertson-Walker geometry of relevance to cosmology.

I.2 The gravitational sector of conformal gravity

Conformal gravity has been advanced as a candidate alternate theory of gravity (see e.g. the reviews Mannheim2006 ; Mannheim2017 and references therein). The theory is a pure metric theory of gravity that possesses all of the general coordinate invariance and equivalence principle structure of standard gravity while augmenting it with an additional symmetry, local conformal invariance, in which the action is left invariant under local conformal transformations on the metric of the form gμν(x)e2α(x)gμν(x)g_{\mu\nu}(x)\rightarrow e^{2\alpha(x)}g_{\mu\nu}(x) with arbitrary local phase α(x)\alpha(x). Under such a symmetry a gravitational sector action that is to be a polynomial function of the Riemann tensor is uniquely prescribed, and with use of the Gauss-Bonnet theorem is given by (see e.g. Mannheim2006 )

IW=αgd4x(g)1/2CλμνκCλμνκ2αgd4x(g)1/2[RμκRμκ13(Rαα)2].\displaystyle I_{\rm W}=-\alpha_{g}\int d^{4}x\,(-g)^{1/2}C_{\lambda\mu\nu\kappa}C^{\lambda\mu\nu\kappa}\equiv-2\alpha_{g}\int d^{4}x\,(-g)^{1/2}\left[R_{\mu\kappa}R^{\mu\kappa}-\frac{1}{3}(R^{\alpha}_{\phantom{\alpha}\alpha})^{2}\right]. (1.2)

Here αg\alpha_{g} is a dimensionless gravitational coupling constant, and

Cλμνκ=Rλμνκ12(gλνRμκgλκRμνgμνRλκ+gμκRλν)+16Rαα(gλνgμκgλκgμν)\displaystyle C_{\lambda\mu\nu\kappa}=R_{\lambda\mu\nu\kappa}-\frac{1}{2}\left(g_{\lambda\nu}R_{\mu\kappa}-g_{\lambda\kappa}R_{\mu\nu}-g_{\mu\nu}R_{\lambda\kappa}+g_{\mu\kappa}R_{\lambda\nu}\right)+\frac{1}{6}R^{\alpha}_{\phantom{\alpha}\alpha}\left(g_{\lambda\nu}g_{\mu\kappa}-g_{\lambda\kappa}g_{\mu\nu}\right) (1.3)

is the conformal Weyl tensor. (Here and throughout unless indicated to the contrary we follow the notation and conventions of Weinberg1972 .)

With the Weyl action IWI_{\rm W} given in (1.2) being a fourth-order derivative function of the metric, functional variation with respect to the metric gμν(x)g_{\mu\nu}(x) generates fourth-order derivative gravitational equations of motion of the form Mannheim2006

2(g)1/2δIWδgμν=4αgWμν=4αg[2κλCμλνκRκλCμλνκ]=4αg[W(2)μν13W(1)μν]=Tμν,\displaystyle-\frac{2}{(-g)^{1/2}}\frac{\delta I_{\rm W}}{\delta g_{\mu\nu}}=4\alpha_{g}W^{\mu\nu}=4\alpha_{g}\left[2\nabla_{\kappa}\nabla_{\lambda}C^{\mu\lambda\nu\kappa}-R_{\kappa\lambda}C^{\mu\lambda\nu\kappa}\right]=4\alpha_{g}\left[W^{\mu\nu}_{(2)}-\frac{1}{3}W^{\mu\nu}_{(1)}\right]=T^{\mu\nu}, (1.4)

where the functions W(1)μνW^{\mu\nu}_{(1)} and W(2)μνW^{\mu\nu}_{(2)} (respectively associated with the (Rαα)2(R^{\alpha}_{\phantom{\alpha}\alpha})^{2} and RμκRμκR_{\mu\kappa}R^{\mu\kappa} terms in (1.2)) are given by

W(1)μν\displaystyle W^{\mu\nu}_{(1)} =\displaystyle= 2gμνββRαα2νμRαα2RααRμν+12gμν(Rαα)2,\displaystyle 2g^{\mu\nu}\nabla_{\beta}\nabla^{\beta}R^{\alpha}_{\phantom{\alpha}\alpha}-2\nabla^{\nu}\nabla^{\mu}R^{\alpha}_{\phantom{\alpha}\alpha}-2R^{\alpha}_{\phantom{\alpha}\alpha}R^{\mu\nu}+\frac{1}{2}g^{\mu\nu}(R^{\alpha}_{\phantom{\alpha}\alpha})^{2},
W(2)μν\displaystyle W^{\mu\nu}_{(2)} =\displaystyle= 12gμνββRαα+ββRμνβνRμββμRνβ2RμβRβν+12gμνRαβRαβ,\displaystyle\frac{1}{2}g^{\mu\nu}\nabla_{\beta}\nabla^{\beta}R^{\alpha}_{\phantom{\alpha}\alpha}+\nabla_{\beta}\nabla^{\beta}R^{\mu\nu}-\nabla_{\beta}\nabla^{\nu}R^{\mu\beta}-\nabla_{\beta}\nabla^{\mu}R^{\nu\beta}-2R^{\mu\beta}R^{\nu}_{\phantom{\nu}\beta}+\frac{1}{2}g^{\mu\nu}R_{\alpha\beta}R^{\alpha\beta}, (1.5)

and where TμνT^{\mu\nu} is the conformal invariant and thus traceless energy-momentum tensor associated with a conformal matter source. With the action being both general coordinate invariant and conformal invariant WμνW^{\mu\nu} is automatically covariantly conserved and covariantly traceless (i.e., it obeys νWμν=0\nabla_{\nu}W^{\mu\nu}=0, gμνWμν=0g_{\mu\nu}W^{\mu\nu}=0) without the need to impose any gravitational equation of motion. And with IWI_{\rm W} being conformal invariant WμνW_{\mu\nu} transforms as Wμνe2α(x)WμνW_{\mu\nu}\rightarrow e^{-2\alpha(x)}W_{\mu\nu} under a local conformal transformation. In addition the Weyl tensor vanishes in geometries that are conformal to flat, so that in the homogeneous and isotropic conformal to flat background Robertson-Walker and de Sitter geometries of interest to cosmology the background WμνW^{\mu\nu} is zero. Despite this, fluctuations (i.e., cosmological inhomogeneities such as the galaxies and stars of interest to us in this paper) around that background are not conformal to flat, with the fluctuating δWμν\delta W^{\mu\nu} then not being zero.

I.3 The matter sector

Since the light cone is left invariant under a local conformal transformation, in a conformal invariant theory there are no mass scales at the level of the Lagrangian, with all particles being massless. The generation of fermion mass scales in a conformal invariant theory has to be done in the vacuum, and is characterized by an order parameter S(x)S(x) (a vacuum expectation value). For the fermion and scalar field sector we take the matter action to be of the form

IM=d4x(g)1/2[12μSμS112S2Rμμ+λS4+iψ¯γcVcμ(x)[μ+Γμ(x)]ψhSψ¯ψ],\displaystyle I_{M}=-\int d^{4}x(-g)^{1/2}\left[\frac{1}{2}\nabla_{\mu}S\nabla^{\mu}S-\frac{1}{12}S^{2}R^{\mu}_{\phantom{\mu}\mu}+\lambda S^{4}+i\bar{\psi}\gamma^{c}V^{\mu}_{c}(x)[\partial_{\mu}+\Gamma_{\mu}(x)]\psi-hS\bar{\psi}\psi\right], (1.6)

where hh and λ\lambda are dimensionless coupling constants. Here Vcμ(x)V^{\mu}_{c}(x) is a vierbein and Γμ=(1/8)[γa,γb](VνbμVaν+VλbΓνμλVaν)\Gamma_{\mu}=-(1/8)[\gamma_{a},\gamma_{b}](V^{b}_{\nu}\partial_{\mu}V^{a\nu}+V^{b}_{\lambda}\Gamma^{\lambda}_{\phantom{\lambda}\nu\mu}V^{a\nu}) is the spin connection, with the γa\gamma_{a} being a set of fixed axis Dirac gamma matrices and Γνμλ\Gamma^{\lambda}_{\phantom{\lambda}\nu\mu} being the Levi-Civita connection (1/2)gλσ[μgσν+νgσμσgμν](1/2)g^{\lambda\sigma}[\partial_{\mu}g_{\sigma\nu}+\partial_{\nu}g_{\sigma\mu}-\partial_{\sigma}g_{\mu\nu}]. As such, this action is locally conformal invariant under gμν(x)e2α(x)gμν(x)g_{\mu\nu}(x)\rightarrow e^{2\alpha(x)}g_{\mu\nu}(x), Vμa(x)eα(x)Vμa(x)V^{a}_{\mu}(x)\rightarrow e^{\alpha(x)}V^{a}_{\mu}(x), ψ(x)e3α(x)/2ψ(x)\psi(x)\rightarrow e^{-3\alpha(x)/2}\psi(x), S(x)eα(x)S(x)S(x)\rightarrow e^{-\alpha(x)}S(x). Variation of this action with respect to ψ(x)\psi(x) and S(x)S(x) yields the matter field equations of motion

iγcVcμ(x)[μ+Γμ(x)]ψhSψ=0,\displaystyle i\gamma^{c}V^{\mu}_{c}(x)[\partial_{\mu}+\Gamma_{\mu}(x)]\psi-hS\psi=0, (1.7)
μμS+16SRμμ4λS3+hψ¯ψ=0,\displaystyle\nabla_{\mu}\nabla^{\mu}S+\frac{1}{6}SR^{\mu}_{\phantom{\mu}\mu}-4\lambda S^{3}+h\bar{\psi}\psi=0, (1.8)

while variation with respect to the metric yields an energy-momentum tensor of the form (first without use of any matter field equation of motion and then with)

Tμν\displaystyle T^{\mu\nu} =\displaystyle= iψ¯γcVμc(x)[ν+Γν(x)]ψ+23μSνS16gμναSαS13SμνS\displaystyle i\bar{\psi}\gamma^{c}V_{\mu c}(x)[\partial_{\nu}+\Gamma_{\nu}(x)]\psi+\frac{2}{3}\nabla^{\mu}S\nabla^{\nu}S-\frac{1}{6}g^{\mu\nu}\nabla_{\alpha}S\nabla^{\alpha}S-\frac{1}{3}S\nabla^{\mu}\nabla^{\nu}S (1.9)
+\displaystyle+ 13gμνSααS16S2(Rμν12gμνRαα)\displaystyle\frac{1}{3}g^{\mu\nu}S\nabla_{\alpha}\nabla^{\alpha}S-\frac{1}{6}S^{2}\left(R^{\mu\nu}-\frac{1}{2}g^{\mu\nu}R^{\alpha}_{\phantom{\alpha}\alpha}\right)
\displaystyle- gμν[λS4+iψ¯γcVcμ(x)[μ+Γμ(x)]ψhSψ¯ψ]\displaystyle g^{\mu\nu}[\lambda S^{4}+i\bar{\psi}\gamma^{c}V^{\mu}_{c}(x)[\partial_{\mu}+\Gamma_{\mu}(x)]\psi-hS\bar{\psi}\psi]
=\displaystyle= iψ¯γcVμc(x)[ν+Γν(x)]ψ+23μSνS16gμναSαS13SμνS\displaystyle i\bar{\psi}\gamma^{c}V_{\mu c}(x)[\partial_{\nu}+\Gamma_{\nu}(x)]\psi+\frac{2}{3}\nabla^{\mu}S\nabla^{\nu}S-\frac{1}{6}g^{\mu\nu}\nabla_{\alpha}S\nabla^{\alpha}S-\frac{1}{3}S\nabla^{\mu}\nabla^{\nu}S
+\displaystyle+ 13gμνSααS16S2(Rμν12gμνRαα)gμνλS4.\displaystyle\frac{1}{3}g^{\mu\nu}S\nabla_{\alpha}\nabla^{\alpha}S-\frac{1}{6}S^{2}\left(R^{\mu\nu}-\frac{1}{2}g^{\mu\nu}R^{\alpha}_{\phantom{\alpha}\alpha}\right)-g^{\mu\nu}\lambda S^{4}.

Use of the matter field equations of motion then confirms that the matter sector energy-momentum tensor is indeed traceless. Moreover, from the conformal invariance of IMI_{\rm M} it follows that TμνT_{\mu\nu} transforms as Tμνe2α(x)TμνT_{\mu\nu}\rightarrow e^{-2\alpha(x)}T_{\mu\nu} under a local conformal transformation. With WμνW_{\mu\nu} transforming the same way, the 4αgWμν=Tμν4\alpha_{g}W_{\mu\nu}=T_{\mu\nu} equation of motion is conformal invariant. We now proceed to look for static, spherically symmetric solutions to these equations. As such, these selfsame equations could be considered to be microscopic or macroscopic, and so we first explore them as macroscopic equations, and defer a discussion of them as microscopic ones to Sec. VI.

II The Equations of Motion in the Static, Spherically Symmetric Case

For a static, spherically symmetric source the Weyl tensor, and thus Wμν=2κλCμλνκRκλCμλνκW^{\mu\nu}=2\nabla_{\kappa}\nabla_{\lambda}C^{\mu\lambda\nu\kappa}-R_{\kappa\lambda}C^{\mu\lambda\nu\kappa}, do not vanish, and thus in trying to solve the equation Wμν=Tμν/4αgW^{\mu\nu}=T^{\mu\nu}/4\alpha_{g}, we have to deal with a fourth-order differential equation. As noted in Mannheim1989 , it turns out that in the static, spherically symmetric case it is possible to greatly simplify the problem through use of the underlying conformal symmetry that the theory possesses. Specifically, under the general coordinate transformation

ρ=p(r),B(r)=r2b(r)p2(r),A(r)=r2a(r)p2(r)p2(r),\displaystyle\rho=p(r),\qquad B(r)=\frac{r^{2}b(r)}{p^{2}(r)},\qquad A(r)=\frac{r^{2}a(r)p^{\prime 2}(r)}{p^{2}(r)}, (2.1)

with an initially arbitrary function p(r)p(r), the general static, spherically symmetric line element

ds2=b(ρ)dt2a(ρ)dρ2ρ2dθ2ρ2sin2θdϕ2\displaystyle ds^{2}=b(\rho)dt^{2}-a(\rho)d\rho^{2}-\rho^{2}d\theta^{2}-\rho^{2}\sin^{2}\theta d\phi^{2} (2.2)

is brought to the form

ds2=p2(r)r2[B(r)dt2A(r)dr2r2dθ2r2sin2θdϕ2].\displaystyle ds^{2}=\frac{p^{2}(r)}{r^{2}}\left[B(r)dt^{2}-A(r)dr^{2}-r^{2}d\theta^{2}-r^{2}\sin^{2}\theta d\phi^{2}\right]. (2.3)

On now choosing p(r)p(r) according to

1p(r)=rdrr2[a(r)b(r)]1/2,\displaystyle-\frac{1}{p(r)}=\int^{r}\frac{dr}{r^{2}[a(r)b(r)]^{1/2}}, (2.4)

the function A(r)A(r) would then be set equal to 1/B(r)1/B(r), with the line element then being brought to the form

ds2=p2(r)r2[B(r)dt2dr2B(r)r2dθ2r2sin2θdϕ2].\displaystyle ds^{2}=\frac{p^{2}(r)}{r^{2}}\left[B(r)dt^{2}-\frac{dr^{2}}{B(r)}-r^{2}d\theta^{2}-r^{2}\sin^{2}\theta d\phi^{2}\right]. (2.5)

While (2.5) is coordinate equivalent to (2.2), the utility of (2.5) lies in the fact that one of the two functions present in it appears purely as an overall multiplier. Consequently, since both WμνW_{\mu\nu} and TμνT_{\mu\nu} transform the same way under a local conformal transformation the function p(r)/rp(r)/r can be removed from the theory altogether (i.e., gauged away) via a local conformal transformation, with the full kinematic content of the conformal theory then being contained in the line element

ds2=B(r)dt2dr2B(r)r2dθ2r2sin2θdϕ2.\displaystyle ds^{2}=B(r)dt^{2}-\frac{dr^{2}}{B(r)}-r^{2}d\theta^{2}-r^{2}\sin^{2}\theta d\phi^{2}. (2.6)

Given (2.6) we can determine the components of WμνW_{\mu\nu} in a closed form. With the trace condition BW00+(1/B)Wrr+2r2Wθθ=0-BW^{00}+(1/B)W^{rr}+2r^{2}W^{\theta\theta}=0 holding, we only need to express W00W^{00} and WrrW^{rr} in terms of B(r)B(r). In Mannheim1989 it was shown that WrrW^{r}_{{\phantom{r}}r} is given without any approximation at all as

Wrr=13r4(1+y3y′′),\displaystyle W^{r}_{{\phantom{r}}r}=\frac{1}{3r^{4}}(1+y^{3}y^{\prime\prime}), (2.7)

where y2=r2B2rBy^{2}=r^{2}B^{\prime}-2rB. Then in Mannheim1994 it was shown that

3B(W00Wrr)=B′′′′+4rB′′′.\displaystyle\frac{3}{B}\left(W^{0}_{{\phantom{0}}0}-W^{r}_{{\phantom{r}}r}\right)=B^{\prime\prime\prime\prime}+\frac{4}{r}B^{\prime\prime\prime}. (2.8)

(The right-hand side of (2.8) is the radial component of 4B\vec{\nabla}^{4}B.) Both (2.7) and (2.8) are remarkably compact.

In the presence of a source the relevant components of (1.4) take the form

B′′′′+4rB′′′\displaystyle B^{\prime\prime\prime\prime}+\frac{4}{r}B^{\prime\prime\prime} =\displaystyle= 34αgB(T00Trr)=f(r),\displaystyle\frac{3}{4\alpha_{g}B}\left(T^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r}\right)=f(r), (2.9)
13r4(1+y3y′′)=Trr4αg,\displaystyle\frac{1}{3r^{4}}(1+y^{3}y^{\prime\prime})=\frac{T^{r}_{{\phantom{r}}r}}{4\alpha_{g}}, (2.10)

with (2.9) serving to define f(r)f(r).

The solution to (2.9) can be determined in closed form and is given by Mannheim1994

B(r)\displaystyle B(r) =\displaystyle= B0(r)r20r𝑑rr2f(r)16r0r𝑑rr4f(r)12r𝑑rr3f(r)r26r𝑑rrf(r).\displaystyle B_{0}(r)-\frac{r}{2}\int_{0}^{r}dr^{\prime}r^{\prime 2}f(r^{\prime})-\frac{1}{6r}\int_{0}^{r}dr^{\prime}r^{\prime 4}f(r^{\prime})-\frac{1}{2}\int_{r}^{\infty}dr^{\prime}r^{\prime 3}f(r^{\prime})-\frac{r^{2}}{6}\int_{r}^{\infty}dr^{\prime}r^{\prime}f(r^{\prime}). (2.11)

In (2.11) we have included an allowable B0(r)B_{0}(r) contribution that identically satisfies B0′′′′+(4/r)B0′′′=0B_{0}^{\prime\prime\prime\prime}+(4/r)B_{0}^{\prime\prime\prime}=0 at all points rr, i.e., including those where f(r)f(r) is non-zero.

To constrain the B0(r)B_{0}(r) term we note that the f(r)f(r)-dependent term is due to sources that are associated with an expressly non-vanishing WμνW_{\mu\nu} (i.e., inhomogeneities in the cosmological background), while the B0(r)B_{0}(r) term is associated with sources for which WμνW_{\mu\nu} expressly vanishes everywhere (viz. the cosmological background itself). There are two ways in which the fourth-order derivative function WμνW_{\mu\nu} could vanish identically everywhere: the Weyl tensor itself could vanish or the Weyl tensor could obey 2κλCμλνκRκλCμλνκ=02\nabla_{\kappa}\nabla_{\lambda}C^{\mu\lambda\nu\kappa}-R_{\kappa\lambda}C^{\mu\lambda\nu\kappa}=0. With the non-vanishing components of of the Weyl tensor being given by B′′2B/r+2(B1)/r2B^{\prime\prime}-2B^{\prime}/r+2(B-1)/r^{2}, for vanishing CμλνκC^{\mu\lambda\nu\kappa} the solution is given as B0(r)=1+γ0rK0r2B_{0}(r)=1+\gamma_{0}r-K_{0}r^{2} where, as discussed in the appendix, γ0\gamma_{0} and K0K_{0} are universal constants that are associated with the cosmological background. For 2κλCμλνκRκλCμλνκ=02\nabla_{\kappa}\nabla_{\lambda}C^{\mu\lambda\nu\kappa}-R_{\kappa\lambda}C^{\mu\lambda\nu\kappa}=0 the two other solutions required for the fourth-order derivative Wμν=0W_{\mu\nu}=0 equation are given by B0(r)=w0+u0/rB_{0}(r)=w_{0}+u_{0}/r. Since this solution has to hold everywhere a finiteness condition at r=0r=0 excludes the u0/ru_{0}/r term. Hence in the following we set B0(r)=1+w0+γ0rK0r2B_{0}(r)=1+w_{0}+\gamma_{0}r-K_{0}r^{2}. On defining B(r)B^{*}(r) as the f(r)f(r)-dependent component of B(r)B(r) that is given in (2.11), we can express the full B(r)B(r) as B(r)=B0(r)+B(r)B(r)=B_{0}(r)+B^{*}(r).

From (2.11) it follows that (2.10) takes the form (see e.g. the analogous discussion in Horne2016 )

112r4[4+0r𝑑rr2f(r)0r𝑑rr4f(r)(r𝑑rr3f(r)2w02)22γ00r𝑑rr4f(r)]=Trr4αg.\displaystyle\frac{1}{12r^{4}}\left[4+\int_{0}^{r}dr^{\prime}r^{\prime 2}f(r^{\prime})\int_{0}^{r}dr^{\prime}r^{\prime 4}f(r^{\prime})-\left(\int_{r}^{\infty}dr^{\prime}r^{\prime 3}f(r^{\prime})-2w_{0}-2\right)^{2}-2\gamma_{0}\int_{0}^{r}dr^{\prime}r^{\prime 4}f(r^{\prime})\right]=\frac{T^{r}_{{\phantom{r}}r}}{4\alpha_{g}}. (2.12)

On evaluating the non-vanishing components of the Weyl tensor for the B(r)B(r) given in (2.11) we obtain

B′′2rB+2r2(B1)=1r30r𝑑rr4f(r)1r2r𝑑rr3f(r)+2w0r2,\displaystyle B^{\prime\prime}-\frac{2}{r}B^{\prime}+\frac{2}{r^{2}}(B-1)=-\frac{1}{r^{3}}\int_{0}^{r}dr^{\prime}r^{\prime 4}f(r^{\prime})-\frac{1}{r^{2}}\int_{r}^{\infty}dr^{\prime}r^{\prime 3}f(r^{\prime})+\frac{2w_{0}}{r^{2}}, (2.13)

to thus confirm that the non-vanishing of the matter source f(r)f(r) leads to a non-vanishing Weyl tensor.

For the metric given in (2.6) the various components of the Ricci tensor take the form

Rrr\displaystyle R_{rr} =B′′2B+BrB,Rθθ=1+rB+B,Rϕϕ=sin2θRθθ,\displaystyle=\frac{B^{\prime\prime}}{2B}+\frac{B^{\prime}}{rB},\quad R_{\theta\theta}=-1+rB^{\prime}+B,\quad R_{\phi\phi}=\sin^{2}\theta R_{\theta\theta},
Rtt\displaystyle R_{tt} =B′′B2BBr,Rαα=B′′+4Br+2Br22r2.\displaystyle=-\frac{B^{\prime\prime}B}{2}-\frac{B^{\prime}B}{r},\quad R^{\alpha}_{\phantom{\alpha}\alpha}=B^{\prime\prime}+\frac{4B^{\prime}}{r}+\frac{2B}{r^{2}}-\frac{2}{r^{2}}. (2.14)

Thus in the solution given in (2.11) the Ricci scalar is given by

Rαα\displaystyle R^{\alpha}_{\phantom{\alpha}\alpha} =3r0r𝑑rr2f(r)1r2r𝑑rr3f(r)2r𝑑rrf(r)12K0+6γ0r+2w0r2.\displaystyle=-\frac{3}{r}\int_{0}^{r}dr^{\prime}r^{\prime 2}f(r^{\prime})-\frac{1}{r^{2}}\int_{r}^{\infty}dr^{\prime}r^{\prime 3}f(r^{\prime})-2\int_{r}^{\infty}dr^{\prime}r^{\prime}f(r^{\prime})-12K_{0}+\frac{6\gamma_{0}}{r}+\frac{2w_{0}}{r^{2}}. (2.15)

Some simplification of the general B(r)B(r) is possible depending on how the matter sources are distributed. Since f(r)f(r) is produced by inhomogeneities in the cosmological background, f(r)f(r) will get contributions from both nearby matter and distant matter. The nearby matter is associated with any given galaxy of interest and the distant matter will be composed of large structures such as clusters of galaxies. While we will need to take the spiral (i.e., non-spherical) nature of galaxies composed of NN^{*} stars into consideration below, for the moment we shall take the nearby matter to consist of a static, spherically symmetric source of radius R0R_{0} with its center situated at the origin of the coordinate system that we are using, so that its contribution to f(r)f(r) is only non-zero in an r<R0r<R_{0} region. Similarly, we shall take the large scale structure contribution to begin at some cluster of galaxies scale R1R_{1}. There is thus a region R0<r<R1R_{0}<r<R_{1} in which f(r)f(r) is zero. In this region B(r)B(r) takes the form

B(R0<r<R1)\displaystyle B(R_{0}<r<R_{1}) =\displaystyle= 1+w0+γ0rK0r216r0R0𝑑rr4f(r)r20R0𝑑rr2f(r)\displaystyle 1+w_{0}+\gamma_{0}r-K_{0}r^{2}-\frac{1}{6r}\int_{0}^{R_{0}}dr^{\prime}r^{\prime 4}f(r^{\prime})-\frac{r}{2}\int_{0}^{R_{0}}dr^{\prime}r^{\prime 2}f(r^{\prime}) (2.16)
\displaystyle- 12R1𝑑rr3f(r)r26R1𝑑rrf(r).\displaystyle\frac{1}{2}\int_{R_{1}}^{\infty}dr^{\prime}r^{\prime 3}f(r^{\prime})-\frac{r^{2}}{6}\int_{R_{1}}^{\infty}dr^{\prime}r^{\prime}f(r^{\prime}).

On defining

2β\displaystyle-2\beta^{*} =\displaystyle= 160R0𝑑rr4f(r),γ=120R0𝑑rr2f(r),w=12R1𝑑rr3f(r),κ=16R1𝑑rrf(r),\displaystyle-\frac{1}{6}\int_{0}^{R_{0}}dr^{\prime}r^{\prime 4}f(r^{\prime}),\qquad\gamma^{*}=-\frac{1}{2}\int_{0}^{R_{0}}dr^{\prime}r^{\prime 2}f(r^{\prime}),\qquad w^{*}=-\frac{1}{2}\int_{R_{1}}^{\infty}dr^{\prime}r^{\prime 3}f(r^{\prime}),\quad\kappa^{*}=\frac{1}{6}\int_{R_{1}}^{\infty}dr^{\prime}r^{\prime}f(r^{\prime}),
w\displaystyle w =\displaystyle= 1+w0+w,u=2β,v=γ0+γ,K=K0+κ,\displaystyle 1+w_{0}+w^{*},\quad u=-2\beta^{*},\quad v=\gamma_{0}+\gamma^{*},\quad K=K_{0}+\kappa^{*}, (2.17)

we can write (2.16) as

B(R0<r<R1)=wKr2+ur+vr,\displaystyle B(R_{0}<r<R_{1})=w-Kr^{2}+\frac{u}{r}+vr, (2.18)

to thus possess Newtonian, linear, and quadratic potentials. The β\beta^{*} and γ\gamma^{*} terms (per star) are fixed by the behavior of f(r)f(r) in the r<R0r<R_{0} region (and would be replaced by NβN^{*}\beta^{*} and NγN^{*}\gamma^{*} for a galaxy with NN^{*} stars), while the ww^{*} and κ\kappa^{*} terms are fixed by the behavior of f(r)f(r) in the r>R1r>R_{1} region. All four of the β\beta^{*}, γ\gamma^{*}, ww^{*} and κ\kappa^{*} terms would be zero unless f(r)f(r) is non-vanishing in the relevant regions, i.e., only if, as per (2.13), the Weyl tensor is non-vanishing in those regions too. While the quadratic κr2-\kappa^{*}r^{2} term has the same form as a K0r2-K_{0}r^{2} de Sitter term we should note that they are not equivalent as the κr2-\kappa^{*}r^{2} term only has the form that it does in the r<R1r<R_{1} region, while the de Sitter form holds for all rr. For the κr2-\kappa^{*}r^{2} term we must use the full (r2/6)r𝑑rrf(r)-(r^{2}/6)\int_{r}^{\infty}dr^{\prime}r^{\prime}f(r^{\prime}) term given in (2.11) in the r>R1r>R_{1} region. The κr2-\kappa^{*}r^{2} term is associated with a non-vanishing Weyl tensor, while the K0r2-K_{0}r^{2} term is associated with a Weyl tensor that vanishes.

As constructed, the existence of the solution given in (2.18) with its dimensionful uu, vv and KK parameters is indicative of spontaneous symmetry breaking, since the very existence of solutions to a set of equations that have lower symmetry than the equations themselves is characteristic of spontaneous symmetry breaking. Since the underlying theory in the conformal gravity case is conformal invariant (and thus scaleless), the existence of a solution to the theory that then has a scale can only be brought about by spontaneous symmetry breaking in a Hilbert space with a spontaneously broken vacuum. This would then be realized by the S(x)S(x) field provided it is not an elementary field but is instead associated with a c-number expectation value of appropriate q-number fields.

In the exterior R0<r<R1R_{0}<r<R_{1} region inserting (2.18) into (2.10) yields

13r4(1+y3y′′)=13r4[1+3uvw2]=Trr(R0<r<R1)4αg.\displaystyle\frac{1}{3r^{4}}(1+y^{3}y^{\prime\prime})=\frac{1}{3r^{4}}\left[1+3uv-w^{2}\right]=\frac{T^{r}_{{\phantom{r}}r}(R_{0}<r<R_{1})}{4\alpha_{g}}. (2.19)

If we now take TrrT^{r}_{{\phantom{r}}r} to vanish in R0<r<R1R_{0}<r<R_{1} (i.e., for the moment with no matter fields in that region), then from (2.19) we fix

w2=1+3uv,\displaystyle w^{2}=1+3uv, (2.20)

and up to a renaming of the coefficients given in Mannheim1989 we recover the exterior solution given there, viz.

B(R0<r<R1)=(1+3uv)1/2+ur+vrKr2,\displaystyle B(R_{0}<r<R_{1})=(1+3uv)^{1/2}+\frac{u}{r}+vr-Kr^{2}, (2.21)

a solution that may also be found in Riegert1984b . In terms of the parameters given in (2.17) the w2=1+3uvw^{2}=1+3uv relation takes the form (w0+w)2+2w0+2w+6γβ+6γ0β=0(w_{0}+w^{*})^{2}+2w_{0}+2w^{*}+6\gamma^{*}\beta^{*}+6\gamma_{0}\beta^{*}=0, with this relation following from (2.12) in any region in which f(r)f(r) vanishes provided TrrT^{r}_{\phantom{r}r} also vanishes in that region. It was for a metric of the form given in (2.21) that the conformal theory was successfully applied to galactic rotation curves in Mannheim2011 ; Mannheim2012 ; O'Brien2012 , with fixed fitted parameters β=1.48×105cm\beta^{*}=1.48\times 10^{5}~{}{\rm cm} per solar mass, γ=5.42×1041cm1\gamma^{*}=5.42\times 10^{-41}~{}{\rm cm}^{-1} per solar mass, γ0=3.06×1030cm1\gamma_{0}=3.06\times 10^{-30}~{}{\rm cm}^{-1}(viz. explicitly a cosmological scale), and κ=9.54×1054cm2\kappa^{*}=9.54\times 10^{-54}~{}{\rm cm}^{-2} (viz. explicitly a cluster of galaxies scale) for all the galactic rotation curves in the entire 138 galaxy sample, and with the particles orbiting in the galaxies being taken to be geodesic.

The derivation leading to (2.21) presupposed that there was no scalar field present in the exterior R0<r<R1R_{0}<r<R_{1} region. For (2.21) to continue to hold when there is scalar field in that region requires that the behavior of S(r)S(r) in the R0<r<R1R_{0}<r<R_{1} region be such that it does in fact cause both T00TrrT^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r} and TrrT^{r}_{{\phantom{r}}r} to vanish there. We thus now explore the degree to which this is or is not the case.

III Solving the Equations of Motion with a Macroscopic Scalar Field

III.1 The general case

To obtain an exact solution to the equations of motion in the exterior R0<r<R1R_{0}<r<R_{1} region in the case when there are non-vanishing fields in that region, we restrict the source to just a macroscopic scalar field S(x)S(x). Following Brihaye2009 we look to see if we can have a non-trivial scalar field and yet nonetheless still get TμνT_{\mu\nu} to vanish in the exterior R0<r<R1R_{0}<r<R_{1} region. Thus in the static, spherically symmetric case we take S(x)S(x) to depend on the radial coordinate rr. Then, given the metric in (2.6) and a scalar sector of the form that (1.8) and (1.9) would take if the fermion sector is excluded, in the exterior R0<r<R1R_{0}<r<R_{1} region we obtain

T00Trr=B3(SS′′2S2),\displaystyle T^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r}=\frac{B}{3}(SS^{\prime\prime}-2S^{\prime 2}), (3.1)
Trr=B2S2+B6SS+2B3rSS+B6rS2+B6r2S216r2S2λS4,\displaystyle T^{r}_{{\phantom{r}}r}=\frac{B}{2}S^{\prime 2}+\frac{B^{\prime}}{6}SS^{\prime}+\frac{2B}{3r}SS^{\prime}+\frac{B^{\prime}}{6r}S^{2}+\frac{B}{6r^{2}}S^{2}-\frac{1}{6r^{2}}S^{2}-\lambda S^{4}, (3.2)

while the pure scalar sector of (1.8) becomes

BS′′+BS+2BrS+S6(B′′+4Br+2Br22r2)4λS3=0.\displaystyle BS^{\prime\prime}+B^{\prime}S^{\prime}+\frac{2B}{r}S^{\prime}+\frac{S}{6}\left(B^{\prime\prime}+\frac{4B^{\prime}}{r}+\frac{2B}{r^{2}}-\frac{2}{r^{2}}\right)-4\lambda S^{3}=0. (3.3)

To find an exterior solution it was suggested in Brihaye2009 to set T00Trr=0T^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r}=0 footnoteC . This then requires that SS′′=2S2SS^{\prime\prime}=2S^{\prime 2}, with integrals

S=aS2,S(r)=1b+ar,\displaystyle S^{\prime}=-aS^{2},\qquad S(r)=\frac{1}{b+ar}, (3.4)

where aa and bb are integration constants. If we insert (3.3) into (3.2) we obtain

Trr=B4SS′′+B2S2B12SS+B6rSSB′′24S2+B12r2S2112r2S2.\displaystyle T^{r}_{{\phantom{r}}r}=-\frac{B}{4}SS^{\prime\prime}+\frac{B}{2}S^{\prime 2}-\frac{B^{\prime}}{12}SS^{\prime}+\frac{B}{6r}SS^{\prime}-\frac{B^{\prime\prime}}{24}S^{2}+\frac{B}{12r^{2}}S^{2}-\frac{1}{12r^{2}}S^{2}. (3.5)

With (3.4) we can rewrite (3.5) as

Trr=aB12S3aB6rS3B′′24S2+B12r2S2112r2S2.\displaystyle T^{r}_{{\phantom{r}}r}=a\frac{B^{\prime}}{12}S^{3}-a\frac{B}{6r}S^{3}-\frac{B^{\prime\prime}}{24}S^{2}+\frac{B}{12r^{2}}S^{2}-\frac{1}{12r^{2}}S^{2}. (3.6)

Now the intent of Brihaye2009 , Horne2016 and Hobson2021 was to have the vanishing of T00TrrT^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r} cause TrrT^{r}_{{\phantom{r}}r} to vanish as well. As we see from (3.6), for an as yet unspecified form for B(r)B(r) this is not the case (and not even if we were to set a=0a=0). To see why TrrT^{r}_{{\phantom{r}}r} could not vanish even in principle absent further information of some sort, we recall that we had noted above that because of its tracelessness and covariant conservation TμνT_{\mu\nu} and thus equally WμνW_{\mu\nu} obey

(B2B1r)(T00Trr)(ddr+4r)Trr=0,(B2B1r)(W00Wrr)(ddr+4r)Wrr=0.\displaystyle\left(\frac{B^{\prime}}{2B}-\frac{1}{r}\right)(T^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r})-\left(\frac{d}{dr}+\frac{4}{r}\right)T^{r}_{{\phantom{r}}r}=0,\qquad\left(\frac{B^{\prime}}{2B}-\frac{1}{r}\right)(W^{0}_{{\phantom{0}}0}-W^{r}_{{\phantom{r}}r})-\left(\frac{d}{dr}+\frac{4}{r}\right)W^{r}_{{\phantom{r}}r}=0. (3.7)

Thus from (3.7) we see that even if we set T00Trr=0T^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r}=0, then at best we would could only constrain TrrT^{r}_{{\phantom{r}}r} to be of the form A/r4A/r^{4}, where AA is a constant that decouples from (3.7). To see whether or not we could still secure A=0A=0 anyway we need to impose the gravitational equations of motion given in (2.9) and (2.10), as this is the only further information involving TrrT^{r}_{{\phantom{r}}r} that is available. And as we shall, while it actually is then possible to secure A=0A=0 it is not automatic, and one can find solutions in which A0A\neq 0 as well.

III.2 Solving the gravitational equations of motion

To facilitate the evaluation of (3.2) and (3.3) we insert (3.4) into both of these two equations and obtain

Trr=a2B2S4aB6S32aB3rS3+B6rS2+B6r2S216r2S2λS4,\displaystyle T^{r}_{{\phantom{r}}r}=\frac{a^{2}B}{2}S^{4}-\frac{aB^{\prime}}{6}S^{3}-\frac{2aB}{3r}S^{3}+\frac{B^{\prime}}{6r}S^{2}+\frac{B}{6r^{2}}S^{2}-\frac{1}{6r^{2}}S^{2}-\lambda S^{4}, (3.8)
2a2BS4aBS32aBrS3+S26(B′′+4Br+2Br22r2)4λS4=0.\displaystyle 2a^{2}BS^{4}-aB^{\prime}S^{3}-\frac{2aB}{r}S^{3}+\frac{S^{2}}{6}\left(B^{\prime\prime}+\frac{4B^{\prime}}{r}+\frac{2B}{r^{2}}-\frac{2}{r^{2}}\right)-4\lambda S^{4}=0. (3.9)

With S(r)=1/(b+ar)S(r)=1/(b+ar), on multiplying through by S4S^{4} and conveniently by 6r26r^{2} we obtain

6r2S4Trr=3a2r2Bar2BS14aBrS1+rBS2+BS2S26r2λ,\displaystyle 6r^{2}S^{-4}T^{r}_{{\phantom{r}}r}=3a^{2}r^{2}B-ar^{2}B^{\prime}S^{-1}-4aBrS^{-1}+rB^{\prime}S^{-2}+BS^{-2}-S^{-2}-6r^{2}\lambda, (3.10)
12a2r2B6ar2BS112arBS1+S2(r2B′′+4rB+2B2)24r2λ=0.\displaystyle 12a^{2}r^{2}B-6ar^{2}B^{\prime}S^{-1}-12arBS^{-1}+S^{-2}\left(r^{2}B^{\prime\prime}+4rB^{\prime}+2B-2\right)-24r^{2}\lambda=0. (3.11)

On inserting (2.18), on using (2.19) (but without setting Trr=0T^{r}_{{\phantom{r}}r}=0), and on setting B(R0<r<R1)B(R_{0}<r<R_{1}) equal to a for the moment generic w+u/r+vrKr2w+u/r+vr-Kr^{2} (i.e., without regard to (2.16) and (2.17)), for the respective TrrT^{r}_{\phantom{r}r} and scalar field equations we then obtain

8αgS4r2(1+3uvw2)=6r2S4Trr=r2(abv6λ3Kb2a2)+r(2b2v2abw2ab)+wb2b23abu,\displaystyle\frac{8\alpha_{g}S^{-4}}{r^{2}}(1+3uv-w^{2})=6r^{2}S^{-4}T^{r}_{{\phantom{r}}r}=r^{2}\left(-abv-6\lambda-3Kb^{2}-a^{2}\right)+r\left(2b^{2}v-2abw-2ab\right)+wb^{2}-b^{2}-3abu, (3.12)
r2(6abv24λ12Kb2+2a2w2a2)+r(6b2v8abw4ab+6a2u)+2wb22b26abu=0.\displaystyle r^{2}\left(-6abv-24\lambda-12Kb^{2}+2a^{2}w-2a^{2}\right)+r\left(6b^{2}v-8abw-4ab+6a^{2}u\right)+2wb^{2}-2b^{2}-6abu=0. (3.13)

From the need to satisfy the scalar field equation given in (3.13) for all rr it follows that

3abv12λ6Kb2+a2wa2=0,3b2v4abw2ab+3a2u=0,wb2b23abu=0.\displaystyle-3abv-12\lambda-6Kb^{2}+a^{2}w-a^{2}=0,\quad 3b^{2}v-4abw-2ab+3a^{2}u=0,\quad wb^{2}-b^{2}-3abu=0. (3.14)

We note that (3.14) does not force w213uvw^{2}-1-3uv to vanish. On inserting (3.14) into the TrrT^{r}_{\phantom{r}r} equation given in (3.12) we obtain

8αgr2(1+3uvw2)=1(b+ar)4[r22(abva2a2w)+2ar3(bwb3au)].\displaystyle\frac{8\alpha_{g}}{r^{2}}(1+3uv-w^{2})=\frac{1}{(b+ar)^{4}}\left[\frac{r^{2}}{2}\left(abv-a^{2}-a^{2}w\right)+\frac{2ar}{3}\left(bw-b-3au\right)\right]. (3.15)

Noting now that for general aa and bb the two sides of (3.15) have a totally different dependence on rr, from (3.15) we can conclude that both sides of (3.15) must vanish separately, i.e., that TrrT^{r}_{{\phantom{r}}r} and thus WrrW^{r}_{{\phantom{r}}r} must vanish after all. We thus obtain

1+3uvw2=0bwb3au=0,abva2a2w=0,\displaystyle 1+3uv-w^{2}=0\quad bw-b-3au=0,\quad abv-a^{2}-a^{2}w=0, (3.16)

so that now we can set w213uv=0w^{2}-1-3uv=0. The latter two relations given in (3.16) are compatible with those given in (3.14). And thus for general aa and bb there is a consistent overall completely exact solution of the form

w2=1+3uv,b(w1)=3au,b2v=2ab+3a2u,6λ+3Kb2=2a2a2w.\displaystyle w^{2}=1+3uv,\quad b(w-1)=3au,\quad b^{2}v=2ab+3a^{2}u,\quad 6\lambda+3Kb^{2}=-2a^{2}-a^{2}w. (3.17)

(The third equations given in (3.16) and (3.17) are not independent of the first two, but for convenience we list them.) With (3.17) we thus recover the result assumed but not apparently derived in Brihaye2009 ; Horne2016 ; Hobson2021 that w2=1+3uvw^{2}=1+3uv.

III.3 A special case: a=0a=0

If we set a=0a=0 so that S(r)=1/bS(r)=1/b, then as noted in Horne2016 , we obtain a metric with no linear term, viz.

w=1,v=0,2λ+Kb2=0,B(r)=1+urKr2,\displaystyle w=1,\quad v=0,\quad 2\lambda+Kb^{2}=0,\quad B(r)=1+\frac{u}{r}-Kr^{2}, (3.18)

with uu undetermined. The reason why uu is undetermined is because if we insert B=1+u/rB=1+u/r into the form for the Ricci scalar given in (2.14) we would find that it drops out identically (as it would of course have to do since the Schwarzschild solution is Ricci flat).

The emergence of the structure given in (3.18) is familiar. Specifically, in Mannheim1991 it was noted that one can transform the metric B(r)=1/A(r)=w+u/r+vrKr2B(r)=1/A(r)=w+u/r+vr-Kr^{2} with a linear term into a metric of the form b(ρ)=1/a(ρ)=w+u/ρKρ2b(\rho)=1/a(\rho)=w^{\prime}+u^{\prime}/\rho-K^{\prime}\rho^{2} where there is no linear term. The procedure is to set a(r)=1/b(r)a(r)=1/b(r) in (2.4), so that (2.4) then integrates to

p(r)r=11+cr,\displaystyle\frac{p(r)}{r}=\frac{1}{1+cr}, (3.19)

where cc is an integration constant. On setting ρ=p(r)=r/(1+cr)\rho=p(r)=r/(1+cr), i.e., r=ρ/(1cρ)r=\rho/(1-c\rho), (2.5) takes the form of (2.2) with

b(ρ)=1a(ρ)=w3uc+uρ+ρ(v2wc+3uc2)ρ2(K+vc+uc3wc2).\displaystyle b(\rho)=\frac{1}{a(\rho)}=w-3uc+\frac{u}{\rho}+\rho(v-2wc+3uc^{2})-\rho^{2}(K+vc+uc^{3}-wc^{2}). (3.20)

On now setting c=a/bc=a/b, through use of (3.17) we find that (3.20) takes the form

b(ρ)=1a(ρ)=1+uρ+2λρ2b2,\displaystyle b(\rho)=\frac{1}{a(\rho)}=1+\frac{u}{\rho}+\frac{2\lambda\rho^{2}}{b^{2}}, (3.21)

to thus be of the same form as the Schwarzschild de Sitter solution of standard gravity. (However, as we discuss in Sec. V, the allowed range for ρ\rho is constrained and differs from the range for ρ\rho that is required in the Schwarzschild de Sitter solution.) Then, with c=a/bc=a/b the conformal factor in (2.5) has the form p(r)/r=b/(b+ar)p(r)/r=b/(b+ar). Up to an overall constant we recognize this form for p(r)/rp(r)/r as being none other than the form for S(r)S(r) given in (3.4). With the scalar field transforming as eαS=rS/p(r)e^{-\alpha}S=rS/p(r) under a conformal transformation, as noted in Hobson2021 this form for p(r)/rp(r)/r precisely brings S(r)S(r) to the constant value S(r)=1/bS(r)=1/b. We discuss the significance of this result in Sec. V.

In addition, we would like to note that the our initial derivation of the form for the metric given in (2.6) was a purely kinematic derivation, with the A(r)=1/B(r)A(r)=1/B(r) relation holding without needing to impose the gravitational equations of motion. And as such, this relation holds in both of the Tμν=0T_{\mu\nu}=0 and Tμν0T_{\mu\nu}\neq 0 regions. However, the identification of a(ρ)a(\rho) with 1/b(ρ)1/b(\rho) in (3.21) only holds in the Tμν=0T_{\mu\nu}=0 region. And not only that, it only holds for the particular metric that satisfies the exterior region gravitational equations of motion. The a(ρ)=1/b(ρ)a(\rho)=1/b(\rho) relation does not hold in the Tμν0T_{\mu\nu}\neq 0 region even though the A(r)=1/B(r)A(r)=1/B(r) relation does. Consequently, unlike in the Tμν=0T_{\mu\nu}=0 region, there cannot be any conformal equivalence between solutions in the Tμν0T_{\mu\nu}\neq 0 region.

III.4 A special case: b=0b=0

While the solutions described above lead to Trr=0T^{r}_{{\phantom{r}}r}=0, there is another solution in which TrrT^{r}_{{\phantom{r}}r} is not forced to be zero at all. Specifically, if we set b=0b=0 in (3.12) and (3.13) we obtain

8αga4r4r2(1+3uvw2)=6a4r6Trr=r2(6λa2),r2(24λ+2a2w2a2)+6ra2u=0,\displaystyle\frac{8\alpha_{g}a^{4}r^{4}}{r^{2}}(1+3uv-w^{2})=6a^{4}r^{6}T^{r}_{{\phantom{r}}r}=r^{2}\left(-6\lambda-a^{2}\right),\quad r^{2}\left(-24\lambda+2a^{2}w-2a^{2}\right)+6ra^{2}u=0, (3.22)

with solution

u=0,12λ=a2(w1),8αga4(w21)=6λ+a2.\displaystyle u=0,\quad 12\lambda=a^{2}(w-1),\quad 8\alpha_{g}a^{4}(w^{2}-1)=6\lambda+a^{2}. (3.23)

Then, with w+10w+1\neq 0 we obtain

16a2αg(w1)=1,12λ=116αg.\displaystyle 16a^{2}\alpha_{g}(w-1)=1,\quad 12\lambda=\frac{1}{16\alpha_{g}}. (3.24)

As we see, just as a=0a=0 leads to v=0v=0, similarly b=0b=0 leads to u=0u=0. Now while this may not be particularly interesting physically it does show that in principle setting f(r)=0f(r)=0 does not force TrrT^{r}_{{\phantom{r}}r} to be zero, and not even after we impose the equations of motion. As we had noted in (3.7), setting T00Trr=0T^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r}=0 does not fix the coefficient of the 1/r41/r^{4} term in TrrT^{r}_{{\phantom{r}}r}. Similarly, setting W00Wrr=0W^{0}_{{\phantom{0}}0}-W^{r}_{{\phantom{r}}r}=0 does not fix the coefficient of the 1/r41/r^{4} term in WrrW^{r}_{{\phantom{r}}r}. The b=0b=0 solution leads to both WrrW^{r}_{{\phantom{r}}r} and TrrT^{r}_{{\phantom{r}}r} behaving as 1/r41/r^{4}, and neither of them has to separately vanish. There is thus no general theorem that the vanishing of T00TrrT^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r} requires the vanishing of TrrT^{r}_{{\phantom{r}}r}, and not even if we impose the gravitational equations of motion. Having now established the general implications of setting T00Trr=0T^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r}=0, in Sec. IV we will study geodesics in the most interesting case, namely where TrrT^{r}_{{\phantom{r}}r} is zero, i.e., we consider S(r)=1/(b+ar)S(r)=1/(b+ar) with neither aa nor bb set equal to zero. In order to do this we first need to relate aa and bb to the metric coefficients.

III.5 Fixing aa and bb

For the general case given in (3.17) we can determine aa and bb in terms of the parameters that appear in B(r)=w+u/r+vrKr2B(r)=w+u/r+vr-Kr^{2}, parameters that are determined in (2.17) by dynamical relations that do not involve either aa or bb at all. (With T00TrrT^{0}_{{\phantom{0}}0}-T^{r}_{{\phantom{r}}r} vanishing if S(r)=1/(b+ar)S(r)=1/(b+ar), the general solution to W00Wrr=0W^{0}_{{\phantom{0}}0}-W^{r}_{{\phantom{r}}r}=0 is then independent of aa or bb; with aa and bb then only being fixed in terms of (w,u,v,K)(w,u,v,K) by the 4αgWrr=Trr4\alpha_{g}W^{r}_{{\phantom{r}}r}=T^{r}_{{\phantom{r}}r} and scalar field equations.)

From the w2=1+3uvw^{2}=1+3uv and b(w1)=3uab(w-1)=3ua relations given in (3.17) we obtain

ab=vw+1.\displaystyle\frac{a}{b}=\frac{v}{w+1}. (3.25)

Then from the 6λ+3Kb2=2a2a2w6\lambda+3Kb^{2}=-2a^{2}-a^{2}w relation given in (3.17) we obtain

b2(3K+(w+2)v2(w+1)2)+6λ=0.\displaystyle b^{2}\left(3K+\frac{(w+2)v^{2}}{(w+1)^{2}}\right)+6\lambda=0. (3.26)

With the quantity uv=2Nβ(Nγ+γ0)uv=-2N^{*}\beta^{*}(N^{*}\gamma^{*}+\gamma_{0}) for a galaxy with NN^{*} stars being found to be very much less than one in the galactic rotation curves fits given in Mannheim2011 ; Mannheim2012 ; O'Brien2012 , and with ww thus being close to one, we can set

ab=v2,b2=8λv2+4K.\displaystyle\frac{a}{b}=\frac{v}{2},\quad b^{2}=-\frac{8\lambda}{v^{2}+4K}. (3.27)

Consequently, we can write S(r)S(r) as

S(r)\displaystyle S(r) =\displaystyle= (v2+4K)1/22(2λ)1/2(1+vr/2)=(v2+4K)1/22(2λ)1/2(1vr2)\displaystyle\frac{(v^{2}+4K)^{1/2}}{2(-2\lambda)^{1/2}(1+vr/2)}=\frac{(v^{2}+4K)^{1/2}}{2(-2\lambda)^{1/2}}\left(1-\frac{vr}{2}\right) (3.28)
=\displaystyle= [(Nγ+γ0)2+4(κ+K0)]1/22(2λ)1/2(1(Nγ+γ0)r2),\displaystyle\frac{[(N^{*}\gamma^{*}+\gamma_{0})^{2}+4(\kappa^{*}+K_{0})]^{1/2}}{2(-2\lambda)^{1/2}}\left(1-\frac{(N^{*}\gamma^{*}+\gamma_{0})r}{2}\right),

with the last two steps following since vr1vr\ll 1 in the galactic regime of interest. Thus in this regime we can set

rS(r)S(r)=vr2=(Nγ+γ0)r2,\displaystyle\frac{rS^{\prime}(r)}{S(r)}=-\frac{vr}{2}=-\frac{(N^{*}\gamma^{*}+\gamma_{0})r}{2}, (3.29)

a quantity that we shall meet in Sec. IV.1. Thus we see that both for S(r)S(r) and rS(r)/S(r)rS^{\prime}(r)/S(r) the contribution to S(r>R0)S(r>R_{0}) due to the sources in r<R0r<R_{0} is of order NN^{*}. This will prove to be very consequential in the following.

IV Implication for Geodesics

IV.1 The concern raised by Horne

While a test particle action of the standard IT=m𝑑sI_{T}=-m\int ds form is not conformal invariant, one can construct a test particle action that is, viz. Mannheim1993

IT=hS(x)𝑑s,\displaystyle I_{T}=-h\int S(x)ds, (4.1)

where S(x)S(x) is a scalar field. The variation of this action with respect to the particle coordinate xλx^{\lambda} leads to the trajectory Mannheim1993

d2xλds2+Γμνλdxμdsdxνds=1S[gλμ+dxλdsdxμds]Sxμ.\displaystyle\frac{d^{2}x^{\lambda}}{ds^{2}}+\Gamma^{\lambda}_{\mu\nu}\frac{dx^{\mu}}{ds}\frac{dx^{\nu}}{ds}=-\frac{1}{S}\left[g^{\lambda\mu}+\frac{dx^{\lambda}}{ds}\frac{dx^{\mu}}{ds}\right]\frac{\partial S}{\partial x^{\mu}}. (4.2)

This trajectory only reduces to

d2xλds2+Γμνλdxμdsdxνds=0\displaystyle\frac{d^{2}x^{\lambda}}{ds^{2}}+\Gamma^{\lambda}_{\mu\nu}\frac{dx^{\mu}}{ds}\frac{dx^{\nu}}{ds}=0 (4.3)

if S(x)S(x) is constant. In Mannheim2012 it was shown that in a general metric of the form ds2=B(r)dt2A(r)dr2r2dθ2r2sin2θdϕ2ds^{2}=B(r)dt^{2}-A(r)dr^{2}-r^{2}d\theta^{2}-r^{2}\sin^{2}\theta d\phi^{2} trajectories that obey the scalar-field-dependent (4.2) with a radially-dependent S(r)S(r) are of the form

cd2tds2+cBBdtdsdrds\displaystyle c\frac{d^{2}t}{ds^{2}}+\frac{cB^{\prime}}{B}\frac{dt}{ds}\frac{dr}{ds} =\displaystyle= cSSdtdsdrds,\displaystyle-\frac{cS^{\prime}}{S}\frac{dt}{ds}\frac{dr}{ds},
d2rds2+A2A(drds)2rA(dθds)2rsin2θA(dϕds)2+c2B2A(dtds)2\displaystyle\frac{d^{2}r}{ds^{2}}+\frac{A^{\prime}}{2A}\left(\frac{dr}{ds}\right)^{2}-\frac{r}{A}\left(\frac{d\theta}{ds}\right)^{2}-\frac{r{\rm sin}^{2}\theta}{A}\left(\frac{d\phi}{ds}\right)^{2}+\frac{c^{2}B^{\prime}}{2A}\left(\frac{dt}{ds}\right)^{2} =\displaystyle= SASSS(drds)2,\displaystyle-\frac{S^{\prime}}{AS}-\frac{S^{\prime}}{S}\left(\frac{dr}{ds}\right)^{2},
d2θds2+2rdθdsdrdssinθcosθ(dϕds)2\displaystyle\frac{d^{2}\theta}{ds^{2}}+\frac{2}{r}\frac{d\theta}{ds}\frac{dr}{ds}-{\rm sin}\theta{\rm cos}\theta\left(\frac{d\phi}{ds}\right)^{2} =\displaystyle= SSdθdsdrds,\displaystyle-\frac{S^{\prime}}{S}\frac{d\theta}{ds}\frac{dr}{ds},
d2ϕds2+2rdϕdsdrds+2cosθsinθdϕdsdθds\displaystyle\frac{d^{2}\phi}{ds^{2}}+\frac{2}{r}\frac{d\phi}{ds}\frac{dr}{ds}+2\frac{{\rm cos}\theta}{{\rm sin}\theta}\frac{d\phi}{ds}\frac{d\theta}{ds} =\displaystyle= SSdϕdsdrds,\displaystyle-\frac{S^{\prime}}{S}\frac{d\phi}{ds}\frac{dr}{ds}, (4.4)

with the prime denoting differentiation with respect to rr. Thus for circular orbits with dr/ds=0dr/ds=0, dθ/ds=0d\theta/ds=0, dt/ds=1/cdt/ds=1/c, s=cts=ct, θ=π/2\theta=\pi/2 we obtain

r2c2(dϕdt)2=rB2+rSS,\displaystyle\frac{r^{2}}{c^{2}}\left(\frac{d\phi}{dt}\right)^{2}=\frac{rB^{\prime}}{2}+\frac{rS^{\prime}}{S}, (4.5)

regardless of whether or not A=1/BA=1/B footnoteE . These orbits thus depart from the standard (r2/c2)(dϕ/dt)2=rB/2(r^{2}/c^{2})(d\phi/dt)^{2}=rB^{\prime}/2 that would be associated with (4.3). Now the successful fitting to the rotation curves of 138 spiral galaxies using (2.18) with universal parameters and no dark matter (or its 276 free galactic dark matter halo parameters for the 138 galaxy sample) that was reported in Mannheim2012 ; Mannheim2011 ; O'Brien2012 relied heavily on the linear potential term given in (2.18). (Flat rotation curves come about through a balance between the falling u/ru/r and the rising vrvr.) However, the analysis given above indicates that to support (2.18) we need a scalar field S(r)=1/(b+ar)S(r)=1/(b+ar). Thus if we are to use (2.18) we should use (4.5) rather than the standard (r2/c2)(dϕ/dt)2=rB/2(r^{2}/c^{2})(d\phi/dt)^{2}=rB^{\prime}/2 that was used in Mannheim2012 ; Mannheim2011 ; O'Brien2012 . This then calls into question Horne2016 the fitting given in Mannheim2012 ; Mannheim2011 ; O'Brien2012 .

To see the difficulty in detail, we note that from S(r)=1/(b+ar)S(r)=1/(b+ar) and from (3.17) we obtain

rSS=arb+ar,ab=(w1)3u=vw+1.\displaystyle\frac{rS^{\prime}}{S}=-\frac{ar}{b+ar},\quad\frac{a}{b}=\frac{(w-1)}{3u}=\frac{v}{w+1}. (4.6)

Now according to (2.21) we can set B=w+u/r+vrKr2B=w+u/r+vr-Kr^{2} where w=(1+3uv)1/2w=(1+3uv)^{1/2}. For weak gravity the fits of Mannheim2012 ; Mannheim2011 ; O'Brien2012 lead to uv1uv\ll 1 in galactic regions where u/ru/r, vrvr and Kr2-Kr^{2} are all very much smaller than one. Consequently, with w1w\approx 1 and w+1vrw+1\gg vr (4.5) takes the form

r2c2(dϕdt)2=u2r+vr2Kr2vrw+1+vr=u2r+vr2Kr2vr2=u2rKr2.\displaystyle\frac{r^{2}}{c^{2}}\left(\frac{d\phi}{dt}\right)^{2}=-\frac{u}{2r}+\frac{vr}{2}-Kr^{2}-\frac{vr}{w+1+vr}=-\frac{u}{2r}+\frac{vr}{2}-Kr^{2}-\frac{vr}{2}=-\frac{u}{2r}-Kr^{2}. (4.7)

Thus as noted in Horne2016 , the linear term drops out identically. Moreover, if we instead use the solution with a=0a=0 (so that we then could use the standard geodesic given in (4.3)), then according to (4.6) vv is then zero identically. Thus the linear term is lost whether we work with S(r)=1/(b+ar)S(r)=1/(b+ar) or with S(r)=1/bS(r)=1/b. With the linear term being central to the conformal gravity fitting to galactic rotation curves this then is the concern raised by Horne and reemphasized by Hobson and Lasenby.

IV.2 Shortcomings of the macroscopic scalar field case

To address the concern identified in (4.7) we note that the linear potential contribution to the rS/SrS^{\prime}/S term in (4.5) is only able to cancel the linear potential contribution to the rB/2rB^{\prime}/2 term in (4.5) because these two contributions are equal. Now according to (2.17), for a single star in the cosmological background the linear potential vv term is given by γ0+γ\gamma_{0}+\gamma^{*}, and thus for a galaxy with NN^{*} stars the vv term is given by γ0+Nγ\gamma_{0}+N^{*}\gamma^{*}. Hence the effect of the IT=hS(x)𝑑sI_{T}=-h\int S(x)ds test particle action is to not only give the test particle a mass that is as big as both the cosmological contribution and the number of stars in the galaxy around which it is orbiting, by depending on NN^{*} it gives the test particle a mass that even changes with the environment in which it propagates. In fact, if we set S=1/(b+ar)S=1/(b+ar), then from (3.27) and (2.17) we see that

1b=((γ0+Nγ)2+4K0+4κ8λ)1/2.\displaystyle\frac{1}{b}=\left(\frac{(\gamma_{0}+N^{*}\gamma^{*})^{2}+4K_{0}+4\kappa^{*}}{-8\lambda}\right)^{1/2}. (4.8)

Thus even if we ignore the cosmological background, 1/b1/b is given by

1b=((Nγ)2+4κ8λ)1/2,\displaystyle\frac{1}{b}=\left(\frac{(N^{*}\gamma^{*})^{2}+4\kappa^{*}}{-8\lambda}\right)^{1/2}, (4.9)

to not only make the magnitude of S(r)S(r) (whether it be 1/(b+ar)1/(b+ar) or 1/b1/b) be of the order of the number of stars NN^{*} located in the r<R0r<R_{0} region, but to even be affected by the contribution of clusters of galaxies in the r>R1r>R_{1} region.

Since the analysis given above would lead to a galactic contribution of order NN^{*} and a cosmological contribution of order γ0\gamma_{0} to the test particle orbit we have to conclude that if there is a macroscopic scalar field, the test particle action IT=S(x)𝑑sI_{T}=-\int S(x)ds could not describe the mass generation mechanism for the test particle. Rather, just like the way e𝑑xλAλe\int dx^{\lambda}A_{\lambda} describes a particle propagating in an external long range electromagnetic field, the IT=S(x)𝑑sI_{T}=-\int S(x)ds action describes a particle propagating in a macroscopic long range scalar field, should one exist. For mass generation we thus have to look elsewhere, something we will do in Sec. VI below.

IV.3 Inconsistencies in the pure scalar field calculation

In regard to the analysis that led us to (4.7), we note that is actually internally inconsistent. Specifically, in order for the test particle and the scalar field to couple to each other so as to provide us with the IT=S(x)𝑑sI_{T}=-\int S(x)ds action in the first place, we would need some interaction between them. If we take the fermion to obey the the Yukawa coupled (1.7), then as shown in Mannheim2021 , in the eikonal approximation the fermion will move on the trajectory described by (4.2), viz. the one associated with the test particle action given in (4.1). Thus, the use of (4.2) requires the existence of a long range scalar field that appears in the hSψhS\psi term in the fermion wave equation. However, its presence in the fermion wave equation requires the presence of the hψ¯ψh\bar{\psi}\psi term in the scalar field wave equation given in (1.8). But then we could not use the fermion-independent scalar field wave equation given in (3.3)) that we did use in order to establish that we could set S(r)=1/(a+br)S(r)=1/(a+br) in the first place. Thus if the fermion-scalar coupling is of significance we could not obtain (4.7). Thus we could only recover (4.7) if the coupling term is not significant, in which case the rB/2rB^{\prime}/2 term would then dominate in (4.5) and the geodesics would be standard.

A further shortcoming of the analysis stems from the fact that the test particle of relevance to galactic rotation curves is itself a star. Thus there will not just be a coupling of the form hSψ¯ψhS\bar{\psi}\psi for the test particle, the test particle will have energy and momentum of its own. Thus, as per (1.9), it will contribute a term of the form

Tμν=iψ¯γcVμc(x)[ν+Γν(x)]ψ\displaystyle T^{\mu\nu}=i\bar{\psi}\gamma^{c}V_{\mu c}(x)[\partial_{\nu}+\Gamma_{\nu}(x)]\psi (4.10)

in the R0<r<R1R_{0}<r<R_{1} region that we are considering. So again the analysis of just a scalar field S(x)S(x) in this region is invalid. Now while this particular modification to the analysis would appear be overwhelmed by the NγN^{*}\gamma^{*} and γ0\gamma_{0} contributions to B(r)B(r), the fermion obeys the Yukawa coupled (1.7), and with SS being of order NN^{*}, solutions to the Dirac equation would be of order NN^{*} too.

Because of its spherical symmetry the analysis that leads to (4.7) only applies to a spherical matter source such a single star and not to a spiral galaxy. And this creates an additional difficulty. Specifically, absent any long range scalar field, the standard approach to weak gravity is to first treat each star as an isolated spherical system and solve for the metric exterior to it. Then, all the ensuing potentials are added up using linear superposition. However, if there is a scalar field in the exterior region we could not use linear superposition unless each star were to have its own independent scalar field. But there is only one scalar field, just as there is only one gravitational field. Thus for more than one star we would need to solve for a scalar field in the presence of many sources, sources that for a spiral galaxy are not distributed spherically. This is not the calculation provided in Sec. III. In Sec. V we discuss this many-source issue in the weak gravity limit.

An even more problematic concern in having a long range, massless scalar field may be seen by considering a massless photon propagating in the r>R0r>R_{0} region. Taking the only source to be the f(r)f(r) source in the r<R0r<R_{0} region, then since the photon does not couple to the scalar field it will travel on a null geodesic associated with B(r)B^{*}(r). Moreover, a massless fermion would equally travel on the same null geodesic. Then if we give the fermion a mass, its trajectory would not be modified very much. Moreover, if we take the test particle to be a star, compared with a whole galaxy (let alone the cosmology of the entire Universe), then to one part in 101110^{11} its orbit should not differ from that of an SS-independent geodesic. (It was precisely this argument that was presented in Mannheim2012 ). Thus whatever is to happen to the test particle we should not expect it to be an effect of order NN^{*}. So again, a macroscopic scalar field could not describe mass generation. Below in Sec. VI we show that an effect of order NN^{*} does not appear if the mass generation is microscopic, since then S(x)S(x) is of order the mass of the test particle.

Moreover, if the fermion were to have a mass such that its coupling to the geometry would be of the form IT=m𝑑sI_{T}=-m\int ds, it would follow the standard geodesics and give rise to orbital velocities of the form r2(dϕ/dt)2/c2=rB/2r^{2}(d\phi/dt)^{2}/c^{2}=rB^{\prime}/2. (If S(x)S(x) is equal to a constant mm there would be no rS(r)/S(r)rS^{\prime}(r)/S(r) term in (4.5).) If the test particle is to get its mass via IT=S(x)𝑑sI_{T}=-\int S(x)ds instead, we should not expect this to lead to a radical departure from r2(dϕ/dt)2/c2=rB/2r^{2}(d\phi/dt)^{2}/c^{2}=rB^{\prime}/2. And yet if there is a macroscopic mass-generating field it does. Hence, there must be no long range, macroscopic scalar field to which the test particle could couple to in the first place, and we must thus dispense with such long range mass-generating scalar fields. Thus we must also dispense with a test particle action of the form IT=S(x)𝑑sI_{T}=-\int S(x)ds and replace it by IT=m𝑑sI_{T}=-m\int ds. While this latter action is not conformal invariant, nonetheless in Sec. VI we shall show that in fact this is the test particle action that actually is relevant, as it is associated with the dynamical symmetry breaking mean-field approximation to a four-fermion theory as associated with critical scaling and anomalous dimensions, with the change in vacuum from normal to broken producing a mean-field approximation that is not conformal invariant.

V Conformal Transformation of the Brihaye-Verbin Solution

In regard to long range, macroscopic scalar fields there is one other issue we wish to address, namely the degree to which the S(r)=1/(b+ar)S(r)=1/(b+ar) and S=1/bS=1/b scalar field solutions can be related. As well as study the solution with S(r)=1/(b+ar)S(r)=1/(b+ar), we can follow Horne2016 and Hobson2021 and study the solution in which a=0a=0. To make contact with this solution we can make coordinate and conformal transformations on the B(r)=w+u/r+vrKr2B(r)=w+u/r+vr-Kr^{2} and S(r)=1/(b+ar)S(r)=1/(b+ar) solution that we obtained above. Thus if we now make the conformal and coordinate transformations described in Sec. III.3, then with r=ρ/(1cρ)r=\rho/(1-c\rho), ρ0=R0/(1+cR0)\rho_{0}=R_{0}/(1+cR_{0}), ρ1=R1/(1+cR1)\rho_{1}=R_{1}/(1+cR_{1}) and c=a/bc=a/b we obtain

b(ρ0<ρ<ρ1)=1a(ρ0<ρ<ρ1)=1+uρ+2λρ2b2.\displaystyle b(\rho_{0}<\rho<\rho_{1})=\frac{1}{a(\rho_{0}<\rho<\rho_{1})}=1+\frac{u}{\rho}+\frac{2\lambda\rho^{2}}{b^{2}}. (5.1)

Thus, as shown directly in Horne2016 by solving the equations of motion with S(r)=1/bS(r)=1/b, the linear term drops out identically.

While this metric has the same generic from as a Schwarzschild de Sitter metric, it only does so in the R0<r<R1R_{0}<r<R_{1} region. In the r>R1r>R_{1} region (viz. ρ>R1/(1+cR1)=ρ1\rho>R_{1}/(1+cR_{1})=\rho_{1}) we would be sensitive to the 12r𝑑rr3f(r)(r2/6)r𝑑rrf(r)-\tfrac{1}{2}\int_{r}^{\infty}dr^{\prime}r^{\prime 3}f(r^{\prime})-(r^{2}/6)\int_{r}^{\infty}dr^{\prime}r^{\prime}f(r^{\prime}) term in (2.11), and so in r>R1r>R_{1} the same coordinate and conformal transformations would lead us to

b(ρ0<ρ)\displaystyle b(\rho_{0}<\rho) =\displaystyle= (1cρ)2B(R0<r)=1+uρ+2λρ2b2(1cρ)22ρ/(1cρ)𝑑rr3f(r)\displaystyle(1-c\rho)^{2}B(R_{0}<r)=1+\frac{u}{\rho}+\frac{2\lambda\rho^{2}}{b^{2}}-\frac{(1-c\rho)^{2}}{2}\int_{\rho/(1-c\rho)}^{\infty}dr^{\prime}r^{\prime 3}f(r^{\prime}) (5.2)
\displaystyle- ρ26ρ/(1cρ)𝑑rrf(r)\displaystyle\frac{\rho^{2}}{6}\int_{\rho/(1-c\rho)}^{\infty}dr^{\prime}r^{\prime}f(r^{\prime})

instead. Since b(ρ)b(\rho) would have to behave as b(ρ)=1+u/ρ+2λρ2/b2b(\rho)=1+u/\rho+2\lambda\rho^{2}/b^{2} for all allowed values of ρ\rho in order to be a Schwarzschild de Sitter metric, the metric given in (5.2) is not of this form and can thus not be identified as a Schwarzschild de Sitter metric.

Now we had noted that the exterior S(r)=1/(b+ar)S(r)=1/(b+ar) and S(r)=1/bS(r)=1/b solutions were conformally and coordinate equivalent. However, we now note that the r=ρ/(1cρ)r=\rho/(1-c\rho) transformation is singular at ρ=1/c\rho=1/c unless cc is negative, in which case the ρ/r=p(r)/r=1/(1+cr)\rho/r=p(r)/r=1/(1+cr) conformal transformation is singular (p(r)p(r) being given in (2.4)). Moreover, while r=0r=0 and ρ=0\rho=0 coincide, if c>0c>0 then rr\rightarrow\infty at ρ=1/c\rho=1/c, while ρ=r/(1+cr)\rho=r/(1+cr) is bounded between 0 and 1/c1/c. Now a reader might initially object that because of the Kr2-Kr^{2} term the range for rr actually is bounded. However that is only the case if KK is positive (though even then the bound is of order r=1/K1/2r=1/K^{1/2}, ρ=1/(c+K1/2)\rho=1/(c+K^{1/2}), i.e., not ρ=1/c\rho=1/c), and the transformation between the S(r)=1/(b+ar)S(r)=1/(b+ar) and the S(r)=1/bS(r)=1/b sectors would anyway be the same even if KK is negative or zero, i.e., even if rr is unbounded. With ρ=r/(1+cr)\rho=r/(1+cr) and r=ρ/(1cρ)r=\rho/(1-c\rho) we see that rr is bounded if c<0c<0 while ρ\rho is bounded if c>0c>0.

Now a pure Schwarzschild geometry (viz. (5.1) with λ=0\lambda=0) would require the range of the radial coordinate to be unbounded, and yet there is no choice for the sign of cc for which rr and ρ\rho could both be unbounded. Thus in this regard too we cannot regard (5.1) with λ=0\lambda=0 as being a pure Schwarzschild metric. Moreover, as had been noted in Mannheim1991 , for either sign of cc one of the rρ/(1cρ)r\rightarrow\rho/(1-c\rho) and S(1+cr)SS\rightarrow(1+cr)S transformations is singular, with the S(r)=1/(b+ar)S(r)=1/(b+ar) and S(r)=1/bS(r)=1/b solutions belonging to conformal sectors that are separate and distinct from one another. Moreover, since the fits of Mannheim2012 ; Mannheim2011 ; O'Brien2012 yield uu negative and vv positive, the fits give c=a/b>0c=a/b>0. Then, since we want to work in a non-compact space in which the range of the radial coordinate is infinite (as must be the case if the global cosmological spatial three-curvature kk is negative, with it being shown in Mannheim1989 ; Mannheim2006 and the appendix that in conformal cosmology kk is negative and given by the negative k=γ02/4K0k=-\gamma_{0}^{2}/4-K_{0}), we must thus work in the sector in which S(r)=1/(b+ar)S(r)=1/(b+ar). I.e., we must work in the sector in which there is a linear potential term vrvr, and thus use B(R0<r)B(R_{0}<r) rather than b(ρ0<ρ)b(\rho_{0}<\rho) for galactic rotation curve fits (regardless in fact of whether or not we are working with a long range scalar field). The B(R0<r)B(R_{0}<r) sector is actually the most general sector of the theory since in it both the second- and fourth-moment integrals of the local inhomogeneous source f(r)f(r) couple to the metric as per (2.17). Thus before making any transformation (either with or without any scalar field S(x)S(x)) we should start with the most general form for the metric, viz. the most general solution to Wμν=0W_{\mu\nu}=0 as given in (2.16).

Now in making the transformation from the metric with a vrvr term to the one without we relied on radial symmetry. However in a spiral galaxy there are NN^{*} identical sources and they are not distributed spherically. For the galaxy the weak gravity metric due to localized sources (each with an f(r)f(r) that is only nonzero in r<R0ir<R_{0}^{i}) is given by

B(r)=i(w+u|rri|+v|rri|).\displaystyle B(r)=\sum_{i}\left(w+\frac{u}{|\vec{r}-\vec{r}_{i}|}+v|\vec{r}-\vec{r}_{i}|\right). (5.3)

Now unlike the single source case this time B(r)B(r) depends not just on the rrir-r_{i} radial magnitude but also on the angle between r\vec{r} and ri\vec{r}_{i} as the galactic disk is not spherically symmetric, with |rri|=(r2+ri22rricosϕi)1/2|\vec{r}-\vec{r}_{i}|=(r^{2}+r_{i}^{2}-2rr_{i}\cos\phi_{i})^{1/2}. Thus on transforming to b(ρ)b(\rho) we obtain

b(ρ)=(1cρ)2B(r)=(1cρ)2i(w+uXi+vXi),\displaystyle b(\rho)=(1-c\rho)^{2}B(r)=(1-c\rho)^{2}\sum_{i}\left(w+\frac{u}{X_{i}}+vX_{i}\right), (5.4)

where

Xi=[ρ2(1cρi)2+ρi2(1cρ)22ρρi(1cρ)(1cρi)cosϕi]1/2(1cρ)(1cρi).\displaystyle X_{i}=\frac{[\rho^{2}(1-c\rho_{i})^{2}+\rho_{i}^{2}(1-c\rho)^{2}-2\rho\rho_{i}(1-c\rho)(1-c\rho_{i})\cos\phi_{i}]^{1/2}}{(1-c\rho)(1-c\rho_{i})}. (5.5)

As we see, the vv-dependent term has not in fact been cancelled, since we are only allowed one overall coordinate transformation on all the coordinates and not separate ones for each individual star. Thus as we see, while for a single star we could remove the term that is linear in ρ\rho, by coordinate and conformal transformations we could not do so for an entire spiral galaxy, as we are only allowed one overall coordinate transformation and one overall conformal transformation.

Finally, we should note that even if we use the metric given (5.1) and set S=1/bS=1/b, then while the constancy of SS now leads to SS-independent geodesic test particle motion, according to (4.8) the mass of the test particle would still be of order NN^{*}, and thus still be unacceptable.

VI Dynamical Symmetry Breaking and a Microscopic Scalar Field

To conclude this paper we turn to the role played by scalar fields in microscopic, quantum-field-theoretic mass generation. Such mass generation has two key features that distinguish it from the macroscopic study that we presented above. The first is that relevant scalar fields only have a spatial variation within the particles to which they give mass. And the second is that in the broken vacuum the action is no longer conformal invariant.

For the first issue we note that prior to the early universe electroweak symmetry breaking phase transition the gravitational source consists of massless fermions and massless gauge bosons. Together they provide the radiation fluid that leads to the background cosmological Robertson-Walker geometry. With the fermions being massless they cannot come to rest and form bound systems such as stars, and they thus have to have wave functions that fill all space. (With k=0k=0 these modes are massless plane waves, for k<0k<0 they are associated Legendre functions, and for k>0k>0 they are Gegenbauer polynomials.) Following the phase transition that occurs as the Universe cools down the fermions then do acquire masses by the electroweak symmetry breaking mechanism. Our actual interest is in this breaking being done by fermion bilinears (see Mannheim2017 for a detailed analysis). For the Dirac equation the vacuum |Ω|\Omega\rangle consists of a filled negative energy sea so that S(x)=Ω|ψ¯ψ|ΩS(x)=\langle\Omega|\bar{\psi}\psi|\Omega\rangle is a constant. As shown in the left panel in Fig. 1, in this vacuum the negative energy modes still permeate all space only now as massive plane waves or their k0k\neq 0 analogs. With Ω|ψ¯ψ|Ω\langle\Omega|\bar{\psi}\psi|\Omega\rangle being a constant it just contributes an additional constant value to the cosmological constant, and at this point the geometry is still homogeneous and isotropic.

Refer to caption
Figure 1: undistorted negative energy states, bound first positive energy state, distorted states

Given this vacuum structure it would thus be expected that the lowest lying positive energy mode would also be a massive plane wave, and that the geometry would remain homogeneous and isotropic with a vanishing Weyl tensor. However, it turns out that there is a positive energy state that lies lower in energy than a massive plane wave. Specifically, as noted in Dashen1974 ; Bardeen1975 ; Dashen1975 , in this state, known as a coherent state |C|C\rangle, the scalar field takes a spatially-dependent expectation value S(x)=C|ψ¯(x)ψ(x)|CS(x)=\langle C|\bar{\psi}(x)\psi(x)|C\rangle. As shown in the middle panel in Fig. 1 with m(x)m(x) denoting S(x)S(x) this expectation value is constant everywhere except in a small region where it forms a potential well. Now on its own this state is not stable. However, in the well that is formed one can bind the lowest lying positive energy fermion with wave function u0(x)=x|b|Cu_{0}(x)=\langle x|b^{\dagger}|C\rangle, with the entire configuration (coherent state plus fermion) being stable, where the creation operator bb^{\dagger} is the conjugate of the annihilation operator bb that effects b|C=0b|C\rangle=0. As shown in the right panel in Fig. 1 the negative energy wave functions are also slightly modified in this region. However, they still take their prior plane wave form asymptotically. In the well in which it is bound the positive energy fermion wave function has both a height and a width, just as needed in order to support both of the moment integrals given in (2.17), with the second moment being due to the height [f(r)f(r) can contain a delta function], and the fourth moment being due to the width [f(r)f(r) needing to contain a derivative of a delta function]. The width can be understood as the radius of the fermion, and since there is a non-vanishing height the contribution of the fermion to the second-moment integral, and thus to the parameter vv, cannot be zero. The dynamically induced scalar field thus plays two roles: it both generates fermion masses and provides a mechanism for fermions to localize. These localized fermions thus provide inhomogeneities in the homogeneous cosmological background, with this localization leading to a nonvanishing Weyl tensor. With the scalar field approaching its asymptotic background value exponentially fast Dashen1975 ; Mannheim1976 , it is only varying within the localized source, with its contribution to the energy-momentum tensor outside of the source being in the form of a cosmological constant. In consequence, outside the extended source the geometry receives no contributions from any spatial variation in S(x)=C|ψ¯(x)ψ(x)|CS(x)=\langle C|\bar{\psi}(x)\psi(x)|C\rangle.

In regard to the second feature, we recall the familiar double-well potential V(ϕ)=μ2ϕ2/2+λϕ4/4V(\phi)=-\mu^{2}\phi^{2}/2+\lambda\phi^{4}/4, where ϕ\phi is a quantum field. In the tree approximation ϕ\phi acquires a vacuum expectation value S(x)=Ω|ϕ|Ω=μ/λ1/2S(x)=\langle\Omega|\phi|\Omega\rangle=\mu/\lambda^{1/2} at the minimum of the potential. To expand around this minimum we set ϕ=S(x)+ϕ~\phi=S(x)+\tilde{\phi} where ϕ~\tilde{\phi} is the quantum fluctuation. The fluctuation potential then takes the form V(ϕ~)=μ4/4λ+μ2ϕ~2+λ1/2μϕ~3+λϕ~4/4V(\tilde{\phi})=-\mu^{4}/4\lambda+\mu^{2}\tilde{\phi}^{2}+\lambda^{1/2}\mu\tilde{\phi}^{3}+\lambda\tilde{\phi}^{4}/4. Consequently, the initial ϕϕ\phi\rightarrow-\phi symmetry of V(ϕ)V(\phi) is lost with V(ϕ~)V(\tilde{\phi}) not having any ϕ~ϕ~\tilde{\phi}\rightarrow-\tilde{\phi} symmetry.

This same pattern is repeated for fermions and is readily seen in the paradigm for dynamical symmetry breaking, namely the chirally-symmetric Nambu-Jona-Lasinio model (NJL) Nambu1961 . As formulated in flat spacetime the model consists of a fermion coupled to a four-fermion interaction with no intrinsic fermion mass term and action

INJL\displaystyle I_{\rm NJL} =\displaystyle= d4x[iψ¯γμμψg2[ψ¯ψ]2g2[ψ¯iγ5ψ]2].\displaystyle\int d^{4}x\left[i\bar{\psi}\gamma^{\mu}\partial_{\mu}\psi-\frac{g}{2}[\bar{\psi}\psi]^{2}-\frac{g}{2}[\bar{\psi}i\gamma_{5}\psi]^{2}\right]. (6.1)

One introduces a mass as a trial parameter and rewrites INJL=IMF+IRII_{\rm NJL}=I_{MF}+I_{RI} in terms of mean field and residual interaction components of the form

IMF\displaystyle I_{MF} =\displaystyle= d4x[iψ¯γμμψmψ¯ψ+m22g],IRI=d4x[g2(ψ¯ψmg)2g2(ψ¯iγ5ψ)2].\displaystyle\int d^{4}x\left[i\bar{\psi}\gamma^{\mu}\partial_{\mu}\psi-m\bar{\psi}\psi+\frac{m^{2}}{2g}\right],\qquad I_{RI}=\int d^{4}x\left[-\frac{g}{2}\left(\bar{\psi}\psi-\frac{m}{g}\right)^{2}-\frac{g}{2}\left(\bar{\psi}i\gamma_{5}\psi\right)^{2}\right]. (6.2)

(This is equivalent to replacing 12[ψ¯ψ]2\tfrac{1}{2}[\bar{\psi}\psi]^{2} by ψ¯ψΩ|ψ¯ψ|Ω\bar{\psi}\psi\langle\Omega|\bar{\psi}\psi|\Omega\rangle.) In the decomposition of the INJLI_{\rm NJL} action into mean-field and residual interaction components we note that neither of these two components is separately chiral symmetric, with it being only their sum that is. The residual interaction thus restores the chiral symmetry that is lost in the mean-field sector, and in doing so it generates massless pseudoscalar Goldstone bosons Nambu1961 (as required by the axial-vector Ward identity) and massive scalar Higgs bosons (see Nambu1961 and e.g. Mannheim2017 ) as dynamical bound states.

In the Hartree-Fock approximation one sets

Ω|[ψ¯ψmg]2|Ω=Ω|[ψ¯ψmg]|Ω2=0,m=gΩ|ψ¯ψ|Ω,Ω|[ψ¯iγ5ψ]2|Ω=Ω|[ψ¯iγ5ψ]|Ω2=0,\displaystyle\langle\Omega|\left[\bar{\psi}\psi-\frac{m}{g}\right]^{2}|\Omega\rangle=\langle\Omega|\left[\bar{\psi}\psi-\frac{m}{g}\right]|\Omega\rangle^{2}=0,\quad m=g\langle\Omega|\bar{\psi}\psi|\Omega\rangle,\quad\langle\Omega|\left[\bar{\psi}i\gamma_{5}\psi\right]^{2}|\Omega\rangle=\langle\Omega|\left[\bar{\psi}i\gamma_{5}\psi\right]|\Omega\rangle^{2}=0, (6.3)

as satisfied by a self-consistent MM according to

MΛ24π2+M34π2ln(Λ2M2)=Mg,\displaystyle-\frac{M\Lambda^{2}}{4\pi^{2}}+\frac{M^{3}}{4\pi^{2}}{\rm ln}\left(\frac{\Lambda^{2}}{M^{2}}\right)=\frac{M}{g}, (6.4)

where Λ\Lambda is a large momentum cutoff. In this approximation the vacuum energy is given by the non-chiral-invariant mean-field sector alone since Ω|IRI|Ω=0\langle\Omega|I_{RI}|\Omega\rangle=0, and is of the form Eguchi1974 ; Mannheim1976

ϵ~(m)\displaystyle\tilde{\epsilon}(m) =\displaystyle= ϵ(m)m22g=id4p(2π)4Trln[m+iϵ]id4p(2π)4Trln[+iϵ]m2g\displaystyle\epsilon(m)-\frac{m^{2}}{2g}=i\int\frac{d^{4}p}{(2\pi)^{4}}{\rm Tr~{}ln}\left[\not{p}-m+i\epsilon\right]-i\int\frac{d^{4}p}{(2\pi)^{4}}{\rm Tr~{}ln}\left[\not{p}+i\epsilon\right]-\frac{m^{2}}{g} (6.5)
=\displaystyle= m416π2ln(Λ2m2)m2M28π2ln(Λ2M2)+m432π2,\displaystyle\frac{m^{4}}{16\pi^{2}}{\rm ln}\left(\frac{\Lambda^{2}}{m^{2}}\right)-\frac{m^{2}M^{2}}{8\pi^{2}}{\rm ln}\left(\frac{\Lambda^{2}}{M^{2}}\right)+\frac{m^{4}}{32\pi^{2}},

where the trace is taken over an occupied negative energy Dirac sea composed of plane waves. The vacuum energy given in (6.5) has the form of a dynamically generated double-well potential with a minimum at m=Mm=M. The effect of the residual interaction is to then radiatively dress the mean-field sector while not changing the Hilbert space built on the vacuum in which Ω|ψ¯ψ|Ω\langle\Omega|\bar{\psi}\psi|\Omega\rangle is non-zero. For our purposes here we note that even though we have generated the fermion mass dynamically through the choice of vacuum, as can be seen from the structure of the mean-field IMFI_{MF}, the orbits associated with the mean-field sector are same as those associated with a fermion with a mechanical mass, i.e., the standard massive particle geodesics that are associated with the test particle action IT=m𝑑sI_{T}=-m\int ds with a constant mass parameter mm.

If we take mm to be a function of the coordinates we have to evaluate Trln[i∂̸m(x)+iϵ]{\rm Tr~{}ln}\left[i\not{\partial}-m(x)+i\epsilon\right] -Trln[i∂̸+iϵ]{\rm Tr~{}ln}\left[i\not{\partial}+i\epsilon\right] (still for the moment in flat spacetime), where now the trace is taken over an occupied negative energy Dirac sea composed of distorted waves. And on doing so obtain an effective action whose leading term is of the logarithmically divergent form Eguchi1974 ; Mannheim1976

IEFF=d4x8π2ln(Λ2M2)[12ημνμm(x)νm(x)+m2(x)M212m4(x)],\displaystyle I_{\rm EFF}=\int\frac{d^{4}x}{8\pi^{2}}{\rm ln}\left(\frac{\Lambda^{2}}{M^{2}}\right)\bigg{[}\frac{1}{2}\eta_{\mu\nu}\partial^{\mu}m(x)\partial^{\nu}m(x)+m^{2}(x)M^{2}-\frac{1}{2}m^{4}(x)\bigg{]}, (6.6)

as written here with metric signature g00>0g_{00}>0. In (6.6) we identify m(x)m(x) as m(x)=gC|ψ¯ψ|Cm(x)=g\langle C|\bar{\psi}\psi|C\rangle, where |C|C\rangle is a coherent state. As constructed, on its own the state |C|C\rangle is not stable. To stabilize it we need to introduce a positive frequency fermion state. Thus we add on to (6.6) a positive frequency one-particle action term

I(1)=d4x[iψ¯(1)γμμψ(1)m(x)ψ¯(1)ψ(1)],\displaystyle I(1)=\int d^{4}x\left[i\bar{\psi}(1)\gamma^{\mu}\partial_{\mu}\psi(1)-m(x)\bar{\psi}(1)\psi(1)\right], (6.7)

with the stationary minimum then being the state u0(x)u_{0}(x) introduced above. With m(x)m(x) only varying within the fermion and with it taking a constant value mm outside, the fermion propagates with mass parameter m(x)=mm(x)=m (i.e., propagation is not affected by internal structure, just as in standard Einstein gravity). Trajectories associated with (6.7) are thus given by setting m(x)=mm(x)=m in (6.7), to thus be standard flat space massive particle geodesics.

Finally, if we evaluate Trln[iγaVaμ(μ+Γμ)m(x)+iϵ]{\rm Tr~{}ln}\left[i\gamma^{a}V^{\mu}_{a}(\partial_{\mu}+\Gamma_{\mu})-m(x)+i\epsilon\right] -Trln[iγaVaμ(μ+Γμ)+iϵ]{\rm Tr~{}ln}\left[i\gamma^{a}V^{\mu}_{a}(\partial_{\mu}+\Gamma_{\mu})+i\epsilon\right] in a background curved space geometry we obtain 'tHooft2010a

IEFF=d4x8π2(g)1/2ln(Λ2M2)[12gμνμm(x)νm(x)+112m2(x)Rαα+m2(x)M212m4(x)].\displaystyle I_{\rm EFF}=\int\frac{d^{4}x}{8\pi^{2}}(-g)^{1/2}{\rm ln}\left(\frac{\Lambda^{2}}{M^{2}}\right)\bigg{[}\frac{1}{2}g_{\mu\nu}\nabla^{\mu}m(x)\nabla^{\nu}m(x)+\frac{1}{12}m^{2}(x)R^{\alpha}_{\phantom{\alpha}\alpha}+m^{2}(x)M^{2}-\frac{1}{2}m^{4}(x)\bigg{]}. (6.8)

With (6.7) being replaced by

I(1)=d4x(g)1/2[iψ¯(1)[iγaVaμ(μ+Γμ)ψ(1)m(x)ψ¯(1)ψ(1)],\displaystyle I(1)=\int d^{4}x(-g)^{1/2}\left[i\bar{\psi}(1)[i\gamma^{a}V^{\mu}_{a}(\partial_{\mu}+\Gamma_{\mu})\psi(1)-m(x)\bar{\psi}(1)\psi(1)\right], (6.9)

and with m(x)m(x) being constant outside the source, trajectories associated with (6.9) (i.e., associated with the positive energy modes that are observed in astrophysics) are thus standard curved space massive particle geodesics.

We recognize (6.8) as being a dynamical symmetry generalization of the (generic) conformally coupled scalar field action given in (1.6), with the c-number m(x)=gC|ψ¯ψ|Cm(x)=g\langle C|\bar{\psi}\psi|C\rangle replacing S(x)S(x) (i.e., m(x)m(x) is not a quantized field that has a canonical conjugate with which it would not commute). Consequently, S(x)S(x) is not an elementary field that appears in a fundamental Lagrangian (i.e., a field that would have a canonical conjugate with which it would not commute). As can be seen from the middle panel in Fig. 1, m(x)m(x) only varies in the same region as the region where the bound positive energy fermion wave function u0(x)u_{0}(x) varies, i.e., solely within the extended fermion source. There is thus no macroscopic scalar field that varies outside the source. Moreover, since the RααR^{\alpha}_{\phantom{\alpha}\alpha} term that is generated in (6.8) is associated with the gμνμm(x)νm(x)g_{\mu\nu}\partial^{\mu}m(x)\partial^{\nu}m(x) term (only the combination of the two of them is locally conformal invariant), it also is only present in the region where m(x)m(x) varies, i.e., solely within the extended source. As we see, through one and the same dynamical symmetry breaking mechanism we generate masses for fermions through Ω|ψ¯ψ|Ω\langle\Omega|\bar{\psi}\psi|\Omega\rangle and localize them through C|ψ¯ψ|C\langle C|\bar{\psi}\psi|C\rangle.

While (6.4), (6.5), (6.6) and and (6.8) involve the cutoff, on using a renormalization group fixed point analysis with subtraction parameter μ2\mu^{2}, in Mannheim1978 ; Mannheim2017 it was shown that via critical scaling with anomalous dimensions we could make ϵ~(m)\tilde{\epsilon}(m) and IEFFI_{\rm EFF} be completely finite footnoteF . Specifically, it was shown that if the dimension dd of θ=ψ¯ψ\theta=\bar{\psi}\psi is reduced from its canonical value of d=3d=3 to d=2d=2 (i.e., θ\theta acquires a dimension d=3+γθd=3+\gamma_{\theta} where γθ=1\gamma_{\theta}=-1) then the vacuum spontaneously breaks and one obtains dynamical Goldstone and Higgs bosons. With the dimension of (ψ¯ψ)2(\bar{\psi}\psi)^{2} being reduced from six to four the four-fermion theory becomes renormalizable, the coupling parameter gg in INJLI_{\rm NJL} action becomes dimensionless, and the theory is now scale invariant, just as required for a matter sector that is to couple to conformal gravity.

At γθ=1\gamma_{\theta}=-1 (6.8) is replaced by Mannheim1978 ; Mannheim2017

IEFF=d4x8π2(g)1/2[m2(x)μ216π2(ln(m2(x)M2)1)+3μ256πm(x)(gκνκm(x)νm(x)+16m2(x)Rαα)],\displaystyle I_{\rm EFF}=\int\frac{d^{4}x}{8\pi^{2}}(-g)^{1/2}\bigg{[}-\frac{m^{2}(x)\mu^{2}}{16\pi^{2}}\left({\rm ln}\left(\frac{m^{2}(x)}{M^{2}}\right)-1\right)+\frac{3\mu}{256\pi m(x)}\left(g_{\kappa\nu}\nabla^{\kappa}m(x)\nabla^{\nu}m(x)+\frac{1}{6}m^{2}(x)R^{\alpha}_{\phantom{\alpha}\alpha}\right)\bigg{]}, (6.10)

with the already curved space, non-conformal invariant, mean-field induced (6.9) remaining as is.

This localization mechanism generates cosmological inhomogeneities in conformal gravity, and allows them to propagate in an exterior geometry in which both μm\nabla_{\mu}m and RααR^{\alpha}_{\phantom{\alpha}\alpha} are zero. In such regions in which f(r)f(r) is zero one thus has B(r)=B0(r)+B(r)B(r)=B_{0}(r)+B^{*}(r), while orbital velocities are given by the standard (r2/c2)(dϕ/dt)2=rB/2(r^{2}/c^{2})(d\phi/dt)^{2}=rB^{\prime}/2 that was used in Mannheim2012 ; Mannheim2011 ; O'Brien2012 . Thus unlike in the case with a long range, massless scalar field, when mass is generated microscopically via dynamical symmetry breaking orbits are standard.

Acknowledgements.
The author wishes to thank Dr. Y. Brihaye, Dr. M. P. Hobson and Dr. A. N. Lasenby for some helpful comments.

Appendix A Connection with Cosmology

In a Robertson-Walker geometry the comoving time line element is given by

ds2=c2dt2a2(t)(1+kρ2/4)2[dρ2+ρ2dθ2+ρ2sin2θdϕ2]\displaystyle ds^{2}=c^{2}dt^{2}-\frac{a^{2}(t)}{(1+k\rho^{2}/4)^{2}}[d\rho^{2}+\rho^{2}d\theta^{2}+\rho^{2}\sin^{2}\theta d\phi^{2}] (A.1)

as written in spatially isotropic coordinates, where a(t)a(t) is the expansion radius and kk is the spatial three-curvature footnoteA . Since the Robertson-Walker metric is conformal to flat, the conformal WμνW_{\mu\nu} tensor vanishes identically. We take the cosmological energy-momentum tensor to consist of a conformally coupled scalar field σ(x)\sigma(x) and a T0μν=(ρ0+p0)UμUν+p0gμνT_{0}^{\mu\nu}=(\rho_{0}+p_{0})U^{\mu}U^{\nu}+p_{0}g^{\mu\nu} perfect fluid that does not couple to the scalar field (as noted in Sec. I, cosmological or grandunified scales are not associated with the mass generation of the standard model elementary particles that make up the T0μνT_{0}^{\mu\nu} perfect fluid). In analog to (1.8) and (1.9) the scalar field wave equation and the full energy-momentum tensor are given by

μμσ+16σRμμ4λ0σ3=0,\displaystyle\nabla_{\mu}\nabla^{\mu}\sigma+\frac{1}{6}\sigma R^{\mu}_{\phantom{\mu}\mu}-4\lambda_{0}\sigma^{3}=0,
Tμν=(ρ0+p0)UμUν+p0gμν+23μσνσ16gμνασασ13σμνσ\displaystyle T^{\mu\nu}=(\rho_{0}+p_{0})U^{\mu}U^{\nu}+p_{0}g^{\mu\nu}+\frac{2}{3}\nabla^{\mu}\sigma\nabla^{\nu}\sigma-\frac{1}{6}g^{\mu\nu}\nabla_{\alpha}\sigma\nabla^{\alpha}\sigma-\frac{1}{3}\sigma\nabla^{\mu}\nabla^{\nu}\sigma
+13gμνσαασ16σ2(Rμν12gμνRαα)gμνλ0σ4,\displaystyle+\frac{1}{3}g^{\mu\nu}\sigma\nabla_{\alpha}\nabla^{\alpha}\sigma-\frac{1}{6}\sigma^{2}\left(R^{\mu\nu}-\frac{1}{2}g^{\mu\nu}R^{\alpha}_{\phantom{\alpha}\alpha}\right)-g^{\mu\nu}\lambda_{0}\sigma^{4}, (A.2)

where the energy density ρ0\rho_{0} and the pressure p0p_{0} are functions of tt, and UμU_{\mu} is a timelike fluid four-velocity that obeys gμνUμUν=1g_{\mu\nu}U^{\mu}U^{\nu}=-1 (in the notation of Weinberg1972 g00g_{00} is negative). With there being no cross term between the σ\sigma sector and the fluid sector, the σ\sigma sector contribution and the perfect fluid contribution to TμνT^{\mu\nu} are both separately covariantly conserved, with the σ\sigma sector containing a cosmological constant term gμνλ0σ4-g^{\mu\nu}\lambda_{0}\sigma^{4}. The full TμνT^{\mu\nu} must vanish identically since WμνW^{\mu\nu} does, and it is able to do so non-trivially Mannheim2006 . With σ(x)\sigma(x) taking a constant value σ0\sigma_{0}, with the fluid being taken to be a radiation fluid prior to the mass generation epoch so that ρ0=3p0=A/a4(t)\rho_{0}=3p_{0}=A/a^{4}(t) (where AA is a constant), and with kk being shown to be negative in the conformal case Mannheim2006 , the exact equations and solution for the expansion radius a(t)a(t) in the vanishing TμνT^{\mu\nu} case are given as Mannheim2006

Aa4+λ0σ04+12a2c2σ02(a˙2+kc2)=0,a˙2+aa¨+kc2+4c2λ0σ02a2=0,\displaystyle\frac{A}{a^{4}}+\lambda_{0}\sigma_{0}^{4}+\frac{1}{2a^{2}c^{2}}\sigma_{0}^{2}(\dot{a}^{2}+kc^{2})=0,\quad\dot{a}^{2}+a\ddot{a}+kc^{2}+4c^{2}\lambda_{0}\sigma_{0}^{2}a^{2}=0,
a2(t)=k(β1)2αkβsinh2(α1/2ct)α,\displaystyle a^{2}(t)=-\frac{k(\beta-1)}{2\alpha}-\frac{k\beta{\rm sinh}^{2}(\alpha^{1/2}ct)}{\alpha}, (A.3)

where the parameters α\alpha and β\beta are defined as

α=2λ0σ02=K0,β=(116Aλ0k2)1/2,\displaystyle\alpha=-2\lambda_{0}\sigma_{0}^{2}=K_{0},\quad\beta=\left(1-\frac{16A\lambda_{0}}{k^{2}}\right)^{1/2}, (A.4)

and where we have introduced K0=2λ0σ02K_{0}=-2\lambda_{0}\sigma_{0}^{2}, a parameter that serves as a de Sitter scale. With this model a non-fine-tuned, dark matter free, fit to the accelerating universe data is obtained Mannheim2006 , with kk indeed being found to be negative.

Appendix B Writing a Robertson-Walker Geometry in Static Coordinates

In order to be able to make contact between cosmology and galaxies, this being of relevance to our analysis above, it is very instructive to rewrite the Robertson-Walker line element in a static, spherically symmetric coordinate system. That we are able to do this in principle is because in a geometry that is homogeneous and isotropic any point in the spacetime can be taken as the origin of coordinates. Thus each time we analyze the rotation curve of any given galaxy we take the center of that particular galaxy to be the origin of coordinates.

To establish the connection that we need we follow Mannheim1989 and make the coordinate transformation

ρ\displaystyle\rho =\displaystyle= 8+4γ0r8(1+γ0rK0r2)1/2(γ02+4K0)r,r=16ρ168γ0ρ+γ02ρ2+4K0ρ2,\displaystyle\frac{8+4\gamma_{0}r-8(1+\gamma_{0}r-K_{0}r^{2})^{1/2}}{(\gamma_{0}^{2}+4K_{0})r},\quad r=\frac{16\rho}{16-8\gamma_{0}\rho+\gamma_{0}^{2}\rho^{2}+4K_{0}\rho^{2}}, (B.1)
ρ1+kρ2/4=ρ1γ0ρ2/16K0ρ2/4=r(1+γ0rK0r2)1/2,\displaystyle\frac{\rho}{1+k\rho^{2}/4}=\frac{\rho}{1-\gamma_{0}\rho^{2}/16-K_{0}\rho^{2}/4}=\frac{r}{(1+\gamma_{0}r-K_{0}r^{2})^{1/2}},

with (A.1) then taking the form

ds2=c2dt2a2(t)(dr2(1+γ0rK0r2)2+r21+γ0rK0r2(dθ2+sin2θdϕ2)),\displaystyle ds^{2}=c^{2}dt^{2}-a^{2}(t)\bigg{(}\frac{dr^{2}}{(1+\gamma_{0}r-K_{0}r^{2})^{2}}+\frac{r^{2}}{1+\gamma_{0}r-K_{0}r^{2}}(d\theta^{2}+\sin^{2}\theta d\phi^{2})\bigg{)}, (B.2)

where γ02/4K0=k-\gamma_{0}^{2}/4-K_{0}=k. With the de Sitter K0K_{0} being positive and with γ0\gamma_{0} being real and non-zero, it follows that the only allowed value of kk for which we can in fact make this transformation at all is negative, just as noted above.

Introducing the conformal time dτ=dt/a(t)d\tau=dt/a(t) enables us to rewrite (B.2) in the form

ds2=a2(τ)[c2dτ2dr2(1+γ0rK0r2)2r21+γ0rK0r2(dθ2+sin2θdϕ2)],\displaystyle ds^{2}=a^{2}(\tau)\bigg{[}c^{2}d\tau^{2}-\frac{dr^{2}}{(1+\gamma_{0}r-K_{0}r^{2})^{2}}-\frac{r^{2}}{1+\gamma_{0}r-K_{0}r^{2}}(d\theta^{2}+\sin^{2}\theta d\phi^{2})\bigg{]}, (B.3)

with (B.3) still describing a Robertson-Walker geometry. Noting now that the Weyl tensor not only vanishes in a Robertson-Walker geometry but also in a conformally transformed one, we now conformally transform (B.3) by multiplying the line element by Ω2(r,τ)\Omega^{2}(r,\tau) where Ω(r,τ)=(1+γ0rK0r2)1/2/a(τ)\Omega(r,\tau)=(1+\gamma_{0}r-K_{0}r^{2})^{1/2}/a(\tau), with (B.3) then taking the form

ds2=(1+γ0rK0r2)c2dτ2dr2(1+γ0rK0r2)r2(dθ2+sin2θdϕ2).\displaystyle ds^{2}=(1+\gamma_{0}r-K_{0}r^{2})c^{2}d\tau^{2}-\frac{dr^{2}}{(1+\gamma_{0}r-K_{0}r^{2})}-r^{2}(d\theta^{2}+\sin^{2}\theta d\phi^{2}). (B.4)

Since this line element is now of the same generic form as the static, spherically symmetric line element ds2=B(r)dt2dr2/B(r)r2dθ2r2sin2θdϕ2ds^{2}=B(r)dt^{2}-dr^{2}/B(r)-r^{2}d\theta^{2}-r^{2}\sin^{2}\theta d\phi^{2} that we met in (2.6) (on replacing τ\tau by tt), we see that a Robertson-Walker geometry is conformally and coordinate equivalent to a static, spherically symmetric geometry, viz. precisely the geometry needed for study of galactic rotation curves, with cosmology precisely providing the B0(r)=1+γ0rK0r2B_{0}(r)=1+\gamma_{0}r-K_{0}r^{2} contribution to B(r)B(r) that we identified in Sec. II.

In making the conformal transformation the time-dependent ρ0=A/a4(τ)\rho_{0}=A/a^{4}(\tau) transforms into the static Ω4ρ0=A/(1+γ0rK0r2)2\Omega^{-4}\rho_{0}=A/(1+\gamma_{0}r-K_{0}r^{2})^{2}, while the time-independent σ0\sigma_{0} transforms into the time-dependent Ω1σ0=σ0a(τ)/(1+γ0rK0r2)1/2\Omega^{-1}\sigma_{0}=\sigma_{0}a(\tau)/(1+\gamma_{0}r-K_{0}r^{2})^{1/2}. To see how these dynamical conditions are obeyed (something that had not been considered in the purely kinematic study given in Mannheim1989 and Mannheim2006 ), we note that for the line element ds2=B(r)dt2dr2/B(r)r2dθ2r2sin2θdϕ2ds^{2}=B(r)dt^{2}-dr^{2}/B(r)-r^{2}d\theta^{2}-r^{2}\sin^{2}\theta d\phi^{2}, the covariant conservation condition for a perfect fluid that only depends on rr is of the form B/B=2p0/(ρ0+p0)B^{\prime}/B=-2p_{0}^{\prime}/(\rho_{0}+p_{0}). (This relation actually holds for the general ds2=B(r)dt2A(r)dr2r2dθ2r2sin2θdϕ2ds^{2}=B(r)dt^{2}-A(r)dr^{2}-r^{2}d\theta^{2}-r^{2}\sin^{2}\theta d\phi^{2}.) Thus for a fluid that obeys ρ0=3p0\rho_{0}=3p_{0}, ρ0\rho_{0} must behave as 1/B21/B^{2}, i.e., precisely as 1/(1+γ0rK0r2)21/(1+\gamma_{0}r-K_{0}r^{2})^{2} just as we had found. For the scalar field, we find that the time-dependent σ(r,τ)=σ0a(τ)/(1+γ0rK0r2)1/2\sigma(r,\tau)=\sigma_{0}a(\tau)/(1+\gamma_{0}r-K_{0}r^{2})^{1/2} will satisfy the scalar field wave equation given in (A.2) as evaluated with the (B.4) metric provided a(τ)a(\tau) obeys

d2a(τ)dτ2+kc2a(τ)+4λ0σ02c2a3(τ)=0.\displaystyle\frac{d^{2}a(\tau)}{d\tau^{2}}+kc^{2}a(\tau)+4\lambda_{0}\sigma_{0}^{2}c^{2}a^{3}(\tau)=0. (B.5)

On converting the second equation in (A.3) from comoving time to conformal time we recover (B.5) just as required.

With particles being massless in the radiation era they propagate along null geodesics, with these geodesics thus being sensitive to both the γ0\gamma_{0} and K0K_{0} terms in (B.4). This dependence will persist after mass generation, and so both of these terms will also contribute to the trajectories followed by massive particles as well. The γ0r\gamma_{0}r term is particularly significant for conformal gravity rotation curve fits, and is found numerically to be one of the largest contributors, with it overwhelming the contribution of any given star to the galactic potential by a factor of order 101110^{11}. Thus as noted in Mannheim2006 , cosmology is central to the systematics of galactic rotation curves.

References