A Survey of Dynamical and Gravitational Lensing Tests in Scale Invariance: The Fall of Dark Matter?

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Article

A Survey of Dynamical and Gravitational Lensing Tests in Scale


Invariance: The Fall of Dark Matter?
André Maeder 1, * , Frédéric Courbin 2,3

1 Department of Astronomy, University of Geneva, Chemin des Maillettes 51, CH-1290 Sauverny, Switzerland
2 ICC-UB Institut de Ciències del Cosmos, Universitat de Barcelona, Martí Franquès, 1, E-08028 Barcelona, Spain;
[email protected]
3 ICREA, Pg. Lluís Companys 23, E-08010 Barcelona, Spain
* Correspondence: [email protected]
arXiv:2410.21379v1 [astro-ph.CO] 28 Oct 2024

Abstract: We first briefly review the adventure of scale invariance in physics, from Galileo Galilei,
Weyl, Einstein, and Feynman to the revival by Dirac (1973) and Canuto et al. (1977). In the way that
the geometry of space–time can be described by the coefficients gµν , a gauging condition given by
a scale factor λ( x µ ) is needed to express the scaling. In general relativity (GR), λ = 1. The “Large
Number Hypothesis” was taken by Dirac and by Canuto et al. to fix λ. The condition that the
macroscopic empty space is scale-invariant was further preferred (Maeder 2017a), the resulting gauge
is also supported by an action principle. Cosmological equations and a modified Newton equation
were then derived. In short, except in extremely low density regions, the scale-invariant effects
are largely dominated by Newtonian effects. However, their cumulative effects may still play a
significant role in cosmic evolution. The theory contains no “adjustment parameter”. In this work,
we gather concrete observational evidence that scale-invariant effects are present and measurable
in astronomical objects spanning a vast range of masses (0.5 M⊙ < M < 1014 M⊙ ) and an equally
impressive range of spatial scales (0.01 pc < r < 1 Gpc). Scale invariance accounts for the observed
excess in velocity in galaxy clusters with respect to the visible mass, the relatively flat/small slope of
rotation curves in local galaxies, the observed steep rotation curves of high-redshift galaxies, and
the excess of velocity in wide binary stars with separations above 3000 kau found in Gaia DR3. Last
but not least, we investigate the effect of scale invariance on gravitational lensing. We show that
scale invariance does not affect the geodesics of light rays as they pass in the vicinity of a massive
galaxy. However, scale-invariant effects do change the inferred mass-to-light ratio of lens galaxies as
Citation: Maeder, A.; Courbin, F. A
compared to GR. As a result, the discrepancies seen in GR between the total lensing mass of galaxies
Survey of Dynamical and
and their stellar mass from photometry may be accounted for. This holds true both for lenses at high
Gravitational Lensing Tests in Scale
Invariance: The Fall of Dark Matter?
redshift like JWST-ER1 and at low redshift like in the SLACS sample. Of note is that none of the
Symmetry 2024, 1, 0. https://doi.org/ above observational tests require dark matter or any adjustable parameter to tweak the theory at any
given mass or spatial scale.

Academic Editor: Firstname Lastname


Keywords: cosmology; dynamics of galaxies; wide binary stars; gravitational lensing
Received: 7 September 2024
Revised: 9 October 2024
Accepted: 14 October 2024
Published: 1. Introduction
The scale-invariant vacuum (SIV) theory rests on the simple and fundamental idea of
examining whether it is possible, and maybe needed, to extend the group of invariances
Copyright: © 2024 by the authors.
subtending gravitation theory, as suggested by Dirac [2]. Dirac emphasized the following:
Licensee MDPI, Basel, Switzerland. “It appears as one of the fundamental principles in Nature that the equations expressing basic laws
This article is an open access article should be invariant under the widest possible group of transformations”. Thus, is there any other
distributed under the terms and fundamental invariance, in addition to the Galilean invariance of Newton’s Mechanics,
conditions of the Creative Commons the Lorentz invariance of special relativity, and the general covariance of general relativity
Attribution (CC BY) license (https:// (GR)? Clearly, if another such invariance exists, its effects must be vanishingly small in
creativecommons.org/licenses/by/ current life as well as in the dynamics of the solar system, and might only apply in rather
4.0/). restrictive and particular conditions.

Symmetry 2024, 1, 0. https://doi.org/10.3390/sym1010000 https://www.mdpi.com/journal/symmetry


Symmetry 2024, 1, 0 2 of 25

The initial developments of theory were performed by Dirac [2] and by Canuto et al.
[3], who proposed a general scale-invariant field equation. Scale invariance means that the
basic equations do not change upon a transformation of the line element of the following
form:
ds′ = λ( x µ ) ds , (1)
where λ( x µ ) is the scale factor, where ds′ refers to GR, and where ds refers to the scale-
invariant space. The properties of the theory have been presented in [1,4] (see also Section 2).
As to the observational tests, several have been successfully studied, with interesting results
[5,6]. Scale-invariant models [1] do predict an accelerated cosmological expansion, not far
but still different from ΛCDM models, with other additional and positive cosmological
tests [1]. Noticeably, the growth of density fluctuations after recombination occurs quite
rapidly without the need for dark matter [7]. This may account for the occurrence of large
galaxies at high redshift, as shown by recent JWST observations. The new equation of
motion [4,24] in scale invariance predicts a tiny additional acceleration that applies in the
velocity direction: it favors accelerated expansion for outward motions, and collapses in a
contraction.
Section 2 presents the historical sequence of developments of the scale-invariant
theory, as well as its most important properties and predictions that are relevant to our tests.
Section 3 focuses on the dynamics of galaxy clusters and Section 4 explores the rotation
curves of galaxies, both at low and high redshifts. The velocity excesses recently found for
wide binary stars in Gaia DR3 are examined in Section 5. Section 6 analyses the differences
seen between the photometric and lensing mass determinations of galaxies. Finally, Section
7 gives the conclusions.

2. The Origin and Properties of the Scale-Invariant Theory


Scale invariance first appeared with Galileo Galilei, who discovered that the natural
physical laws are not scale-invariant, hence pointing to the existence of scaling laws in physics,
e.g., [8]. Indeed, the presence of matter in a system, by fixing some scales, tends to suppress
scale invariance, as pointed out by Feynman [9]. Scale invariance was considered by Weyl
[10] and Eddington [11] in an attempt to account for electromagnetism by a geometrical
property of space–time. Their original proposal was then abandoned because Einstein
[12] had shown that, in this case, the properties of a particle would depend on its past
world line, so that atoms in an electromagnetic field would not show sharp lines. Scale
invariance has often reappeared in theoretical physics, e.g., in Einstein–Cartan gravity with
possible implications for fermionic dark matter and properties of the Higgs boson [13]. The
notion of scale invariance also appears in the study of fractal distributions in physics and
astrophysics. For example, in astrophysics, the initial density fluctuations created by the
inflation has been advocated for; this also appears in the distribution of voids and matter
density in the Universe [14].
The Weyl Integrable Geometry (WIG) was proposed by Dirac (1973) [2] and by Canuto
et al. (1977) [3] (see also Bouvier and Maeder (1978) [15]). The length, ℓ, of a vector
transforms the same way as in Equation (1), and in the transport of a vector from a point
P1 ( x µ ) to a nearby point P2 ( x µ + dx µ ), the length ℓ of a vector is assumed to change by
dℓ = ℓ κν dx ν . The term κν is called the coefficient of metrical connection; it is a fundamental
characteristic of the geometry, alike to gµν (in GR κν = 0). The integrability condition in
WIG makes it so that Einstein’s criticism does not apply: the coefficient κν must be a perfect
differential, ∂µ κ ν = ∂ν κ µ . Thus, the integral of the change in length on a closed loop is zero,
implying that the integration does not depend on the path followed; see, e.g., [17] for more
details.
The field equation in the scale-invariant theory was demonstrated by Canuto et al. [3]
in the line of results by Dirac [2]. More recently, this field equation has been obtained from
an action principle [4]. In the same way as the field equation in GR requires a metric to
determine the geometry of a system, a gauging condition is needed in the scale-invariant
theory to fix the scale factor, λ. Indeed, in this context, it is useful to recall the following
Symmetry 2024, 1, 0 3 of 25

statement by Einstein [18]: “. . . the existence of rigid standard rulers is an assumption suggested
by our approximate experience, assumption which is arbitrary in its principle”. Dirac [2,19], as
well as Canuto et al. [3], used the so-called “Large Number Hypothesis” (LNH) to fix λ.
Another gauging condition was later preferred by Maeder [1]: the macroscopic empty space is
scale-invariant; hence, the name of scale vacuum invariant (SIV) theory is often employed.
This choice is justified, since the usual equation of the state for the vacuum pvac = −ϱvac c2
is precisely the relationship permitting ϱvac to remain constant for an adiabatic expansion
or contraction [20]. This condition leads to some differential equations, the solution of
which gives the scale factor λ = t0 /t [1], where t is the time of cosmological models; this
time is shown, for example, in Equation (2). Subsequently, the search of a minimum of
the action integral for small λ-variations was shown in [4] to lead to the same differential
equations and scale factor as the above-mentioned gauging condition, which thus receives
a strong basis.
General relativity (GR) and Maxwell equations are indeed scale-invariant in the empty
space without charges and currents, an often ignored property. A single atom in the
Universe is likely to not be sufficient to kill scale invariance in an infinite Universe. Maeder
and Gueorguiev [21] have discussed the problem of the physical connections in a scale-
invariant Universe, as well as the occurrence of inflation, with the conclusion that scale
invariance is certainly forbidden in Universe models with a mean density, ϱ, higher than
the critical ϱc = 3H02 /(8πG ). This is in agreement with the conclusions of Galileo Galilei
and Feynman that matter kills scale invariance, with the difference that scale invariance is
an open possibility [21] for Universe models with a mean density lower than the critical
level; this conclusion is in agreement with the cosmological solutions below.
The equations of the cosmological models [1] resulting from the scale-invariant field
equation with the FLWR metric and the mentioned gauging condition lead to [22]
2/3
t3 − Ωm

a(t) = with tin = Ω1/3
m . (2)
1 − Ωm

The cosmological time t varies between tin = Ω1/3 m at the origin, where a(tin ) = 0, and
the present time t0 = 1, where a(t) = 1. Equation (2) confirms the absence of solution for
Ωm ≥ 1, while it predicts possible physical solutions for Ωm < 1. Figure 1 shows that the SIV
effects are drastically reduced for Ωm differing from zero. After a steep decrease, the effects
are smaller and smaller for higher Ωm ; they only totally vanish for Ωm ≥ 1 (for a mean density
≥ 10−29 g cm−3 ). This is the scheme we are proposing for the breakdown of scale invariance
mentioned by Galileo Galilei and Feynman. Since our Universe, according to all current
results (e.g., Planck Collaboration [23]) has a mean density ϱ such that (ϱ/ϱc ) = Ωm ≤ 0.30, it
appears that scale invariance is a possibility. There is a relation between the dimensionless time
scale t and the current time units τ (in Gyr or seconds):

τ − τin t − tin
= , (3)
τ0 − τin t0 − tin

expressing that the age fraction with respect to the present age is the same in both timescales.
This means that
τ
t = tin + (1 − tin ). (4)
τ0
Times t and τ are the same physical time: simply, when the first goes from tin to 1, the
second goes from 0 to 13.8 Gyr. The dynamical tests presented in this work are based on
the equation of motion, derived from the geodesic for weak fields [2,3], also derived from
an action principle by [15]. This equation of motion, corresponding to a modified Newton
equation, is—at the present time—τ0 [4,24]:

d2 r GM (τ ) r ψ0 dr
=− + , with ψ0 = 1 − Ω1/3
m , (5)
dτ 2 r2 r τ0 dτ
Symmetry 2024, 1, 0 4 of 25

where τ0 is the present age of the Universe in current time units. The additional term on the
right of Equation (5) is an acceleration in the direction of the motion, we call it the dynamical
gravity; it is usually very small, since τ0 is very large. This term, proportional to the velocity,
favors collapse during a contraction, and outwards acceleration in an expansion (in fact, a
corresponding acceleration term appears in the cosmological Equations [1]). The parameter
ψ0 applies at the present epoch; for other times, τ, instead of ψ0 , one has

t0 − tin
ψ(τ ) =  . (6)
tin + τ
τ0 ( t0 − tin )
q
ϱc
The ratio of the new term to the Newtonian one behaves like ϱ in systems at
equilibrium [25]. Thus, at the present epoch, the dynamical gravity is generally negligible.
However, cumulative effects may appear, especially at low densities, on long time scales of
the order of H0−1 .

Figure 1. The red curve shows the scale factor λ = tin /t0 at the origin of the Universe (a(tin ) = 0) for
flat models, with k = 0, as a function of the present-time Ωm . The present scale factor at t0 is λ = 1
for any Ωm (dashed black line). Thus, during the evolution of the Universe from the Big Bang to
present time, the value of λ is only vertically moving, for a given Ωm , from the red curve to the black
dashed line. This shows that, for increasing densities, the amplitudes of the variations of the scale
factor λ are very much reduced; tin is given in Equation (2).

From the conservation laws [1], it comes out that the potential Φ = GM/r is conserved
over the evolution of the Universe, while the masses, alike in special relativity, are variable,
scaling here like M ∼ λ−1 ∼ t. These variations are very limited, due to the small range of
λ (Figure 1). As an example, there is a change in mass by a fraction of 0.0025 or 0.0046 over
the last hundred million years for Ωm = 0.30 and 0.05, respectively (new solar models by
Prof. G. Meynet show that the effects on the present solar luminosity are negligible, leading
to solar models deviating by less than 1% from the observed value; see also Section 3 of
[26], which showed that the small changes in the product GM ∼ t surprisingly have no
significant effect on the present solar luminosity). The above considerations are theoretical
ones, suggesting that some effects of scale invariance are possible in our low-density
Universe. The answer belongs to experiments. This is why here we are studying several
observational tests, that were not considered in [5].
The mathematical tools of the scale-invariant theory may look a bit more complex than
those in GR, since the mathematical developments of the scale-invariant theory require,
instead of tensors, the introduction of cotensors, i.e., mathematical tools which, in addition
to the general covariance of GR, also include the scale covariance [2,3,15] (for a brief
summary see also [3,17]). This slightly modifies the Newton Equation (5), as shown nearly
half a century ago [24] (see also [4,25], some conservation laws in [1,3], as well as the
cosmological Equations in [1]). Despite these differences, the scale invariance theory is
Symmetry 2024, 1, 0 5 of 25

in full agreement with a fundamental statement about gravitation by Einstein [18], who
considered that a simple fundamental mathematical principle must determine the equations.
There are interesting similarities and differences between the scale-invariant theory
and the current scalar-tensor theories. The similarity is that the scale-invariant theory
effectively contains a scalar field φ, in addition to the tensor formalism. In general, in
scalar-tensor theories of gravitation [16,27], the scalar field is defined independently, in
order to specifically study some particular physical situations, such as the neighborhood
of black holes or inflation. The specific difference of the scale-invariant theory is that the
definition of the scalar field φ is based on the scale factor λ,

∂ ln λ
φ = ln λ, with κν = − . (7)
∂x ν
The properties of λ are determined by a gauging condition. The choice we made (see
above) is resting on the properties of the macroscopic vacuum and is therefore a universal
function λ = 1/t, independent on the matter properties. Thus, since the macroscopic
vacuum is assumed the same everywhere, so does λ and thus φ. Thus, there is no freedom
to choose a particular field for a specific purpose, and this is a constraining property of the
scale-invariant theory, which is thus different from usual scalar-tensor theories. We note
that this field φ or λ is acting in the Newton-like approximation, as well it may may play
the role of the rolling field in the inflation [21]. Considerable theoretical developments of
scalar-tensor theories have recently have been performed in terms of Lorentz fiber-bundles
by Ikeda et al. [28]. The fiber-bundle locally represents a multidimensional space as a
projection on another globally different topological structure. This framework is providing
a new valuable approach of the cosmological dynamics for the scalar-tensor theories. It has
been applied by [28] both to the inflation and to the later stages of cosmological evolution,
where it introduces some extra terms in the cosmological equations. The study of this new
possible theoretical approach in the scale-invariant context is however beyond the scope
of the present paper. At this stage, we note that that the scale-invariant theory is, like GR,
invariant to Lorentz transformations, since the change in scale applies the same to the time
and spatial length, keeping the velocities constant.

3. The Mass Excess in Clusters of Galaxies


For the derivations and information about the equations used in this work, we refer
the reader to [1,4]. The possibility of dark matter appeared in 1933 with the work of Zwicky
[29]. However, it was not until around 1980 that the hypothesis of dark matter was generally
recognized. Indeed, large excesses of Virial masses in clusters have repeatedly been found
for nearly a century [29–33]. Virial masses in excess with respect to stellar luminosities
have been confirmed, reaching a factor of about 30–50 [34] for large clusters with masses
between 1014 and 1015 M⊙ . This implies that the stellar mass fraction is only of the order
of 0.02–0.03. Today, it appears that dark matter has been progressively upgraded from its
initial status of an hypothesis to that of a material reality.
X-ray observations of the emitting gas allow estimates of the gas mass fraction (typi-
cally 0.10–0.15) in clusters [35–40], which largely dominates over the stellar mass fraction
mentioned above. These two fractions contribute to the baryon mass fraction, f bar , which is
typically around 0.15–0.18, a factor of about 6 smaller than unity; this result is in agreement
with the Planck Collaboration [23] for matter, baryon, and dark matter density parameters:
Ωm = 0.315 ± 0.007, Ωb = 0.050 ± 0.0002, and ΩDM = 0.265 ± 0.002. Noticeably, a relation
between the total dynamical mass of clusters and the baryonic masses of central regions
has also been found [41].
We may attempt a numerical estimate of the relation between masses and velocity dis-
persions in the scale-invariant theory, using the fact that the scale factor, and the quantities
which depend on it, have limited variations and may be considered as adiabatic invariants.
As the Virial theorem does not apply in an expanding system, we start directly from the
Symmetry 2024, 1, 0 6 of 25

equation of motion (5) and consider a galaxy i with a velocity υi that is attracted by another
j-one at time τ and distance rij . Its acceleration at any time τ occurs according to Equation
(5); further,

dυi Gm j (τ ) ψ(τ ) 1
= + υ, with ψ(τ ) = . (8)
dτ rij2 τ0 i Ω1/3
m
+ τ
1−Ω1/3
m
τ0

ψ(τ ) 1 Ω1/3
m
thus = , with τΩ = τ0 . (9)
τ0 τΩ + τ 1 − Ω1/3
m

The first equation is then multiplied by υi = drij /dτ, integrated and summed up on all
the other ( N − 1) m j -masses. As the variations of mass m j (τ ) are very limited, we consider
a mean value p m j (τ0 ) of masses m j at half the age of the Universe; thus, p is given by
p ≈ (1 + Ω1/3
m ) /2. One also has, when summing all the N i-masses,

1 1 G m j (τ0 ) Z
1
2∑ ∑ ∑ rij ∑ [τΩ + τ ] υi2 dτ .
2
υ i = p + (10)
i
2 i j ̸ =i i

The factor 1/2 in front of the double summation is necessary in order to not account for
the same masses twice. We divide the above equation by N. The term on the left becomes
1 2
2 υ . The first term on the right gives the potential energy pqGM/R of the cluster with a
total present mass M and a radius R, with q as a structural factor, depending on the density
profile (q = 3/5 for an homogeneous distribution). The last term is an adiabatic invariant,
for which we take an average over the lifetime of the cluster between τ1 and τ0 , where τ1 is
the formation epoch. We integrate it as follows:

dτ τ + τ0 τ + τ0
Z τ0
f υ2 = f υ2 ln Ω = f υ2 × x (Ωm ), with x (Ωm ) = ln Ω . (11)
τ1 [τΩ + τ ] τΩ + τ1 τΩ + τ1

The factor f accounts for the integration of υ2 over time τ; a value f = 0.5 is likely a
minimum value. The factors x (Ωm ) are given by Equation (11), taking a ratio τ1 /τ0 = 3%,
which corresponds to galaxy formation at a cosmic time of 400 Myr, with the choice of this
initial time not being critical. The values of x (Ωm ) are x (Ωm ) = 0.387, 0.516, 0.735, and
0.950, respectively, for Ωm = 0.30, 0.20, 0.10, and 0.05. The integration constant on the first
term in Equation (10) may be considered as zero, since a zero velocity is an appropriate
value for the end of the collapse. The same is true for the potential, which vanishes at large
distances. Equation (10) can be written as follows:

pq GMSIV
υ2 = 2 + 2 x ( Ωm ) × f υ2 , (12)
R
where MSIV is the present mass in the scale-invariant theory. In the standard case, a value
Mstd would be derived from the same value of the velocity dispersion υ2 by the usual
relation, υ2 = 2 q GMstd /R. From Equation (12), MSIV is evidently smaller than Mstd . The
ratio of the two mass estimates is given by

Mstd p
≈ , (13)
MSIV [1 − 2 f x (Ωm )]

all masses being considered at the same time for not too distant clusters. The same
structural factor q is taken in the two cases. The statistical factors, which account for
the fact that the observed velocities are the radial ones, are the same in the two the-
ories. The factor p has no effect in the term responsible for the divergence. One has
p = 0.835, 0.792, 0.732, and 0.684 for Ωm = 0.30, 0.20, 0.10, and 0.05, respectively.
The above expression shows that the standard mass estimates are generally much
larger than the SIV ones. This is a result of the cumulative contribution over the ages of
Symmetry 2024, 1, 0 7 of 25

Mstd
the additional acceleration in the equation motion. This may explain why ratios M as
SIV
high as 80 [30] or mass–luminosity ratios M/L of about 700 [32] have been obtained in the
Mstd
past. The ratio M is particularly large and often diverging for low values of Ωm . For
SIV
Mstd
Ωm = 0.05 and f = 0.5, the ratio is M = 13.7. For higher f -values, it diverges, since the
SIV
denominator of (13) vanishes. For larger Ωm , e.g., 0.20, the mass ratios are 4.5, 11.1, and
∞ for f =0.8, 0.9, and 1.0. Thus, if scale-invariant effects are present in nature, we expect
that applying standard mass estimates result in large overestimates of the masses of galaxy
clusters.
Within the scale-invariant theory, there is no need to advocate for dark matter to
account for the large observed velocity dispersions in galaxy clusters. This result, joined to
the other ones noted in the introduction and to those further presented in this article, offers
direct support to the scale-invariant theory. Numerical simulations would be very helpful
to better quantify the various effects already suggested by the above analytical approach.

4. Galactic Rotation
4.1. Binet Equation, Angular Momentum, Secular Variations, and Velocities
The concept of dark matter has been greatly promoted [42] by the observations of the
flat, or relatively flat, rotation curves of galaxies over a few tens of kpc (see review by [43]).
The observations much differ from the standard predictions, which—in the outer regions √
of spiral galaxies—predict orbital velocities υ that behave similarly to Keplerian in 1/ r,
where r is the galactocentric radius. This was a severe disagreement with respect to the
predictions of classical mechanics.
This is why it is worth exploring the potentialities of the scale-invariant theory in
this context. The properties of galactic rotation in the outer layers are mainly determined
by those of a two-body system with a central mass and a test particle. The dynamic of
wide binary stars in Section 5.1 is also governed by the same equations; thus, it is worth
briefly considering the problem in detail. These properties were originally derived nearly
half a century ago [24] in the system of the cosmological dimensionless t-units. Here, we
give a demonstration of this in the current τ units (years and seconds), which brings some
changes in the formal writing of equations. From the equation of motion in current units
(Equation 5), the corresponding two equations in plane polar coordinates (r, ϑ ) are

GM ψ dr ψ
r̈ − r ϑ̇2 = − + , and r ϑ̈ + 2ṙ ϑ̇ = r ϑ̇. (14)
r2 τ0 dτ τ0

When not specified, ψ is ψ(τ ), as given by Equation (8); similarly, M is M(τ ). The 2nd
of the above equations has the following for an integral:

ψ 2
r ϑ̇ = L, (15)
τ0

L is a constant in [cm2 rad s−2 ]. This is a generalization of the classical angular momentum
conservation r2 ϑ̇ = const., by mass unit. It shows that r2 ϑ̇ slowly increases with time.
We now express the content of the three above equations into one single relation. With
Equation (15), the derivative of r with respect to ϑ can be written as

dr dr dτ ṙ r2 ψ r ′ L τ0
r′ ≡ = = thus ṙ = . (16)
dϑ dτ dϑ L τ0 r2 ψ

This kind of relation between r ′ and ṙ can be extended to r̈,

L2 r ′′ τ02 2r ′2 L2 τ02 r ′ L
r̈ = − + 2 , (17)
r 4 ψ2 r 5 ψ2 r
Symmetry 2024, 1, 0 8 of 25

since ψ̇ = −ψ2 /τ0 . Expressing the first equation of (14) with (17), we obtain,

L2 τ02 ′′ r ′2 L2 τ 2 G M
 
4 2
r − 2 − 3 20 + 2 = 0 , (18)
r ψ r r ψ r

where—remarkably and fortunately—the last terms in Equations (5) and (17) have simpli-
fied. Now, writing this last equation with u = 1/r is leading simply to,

GM ψ2
u′′ + u = . (19)
L2 τ02

This is the modified form of the well-known Binet equation, describing planetary
trajectories. We will again find it in the study of lensing as Equation (49), also containing
both post-Newtonian and scale-invariant terms. If the 2nd member is equal to 0, the
solution is u = C cos ϑ, the polar equation of a straight line. Now, we search a solution of
GM ψ2
the form u = p + C cos ϑ and obtain p = L2 τ02
[cm−1 ], known as the latus rectum in the
geometrical construction of conics. C is also a constant in [cm−1 ]. Thus, the solution r (ϑ ) is

1 r0
r= or r= , (20)
GM ψ2 1 + e cos ϑ
L2 τ02
+ C cos ϑ

where e is the eccentricity, determined by the initial conditions. It is dimensionless and


therefore a scale-invariant quantity. If the eccentricity is zero, then r0 = 1/p,

C L2 τ02 L2 τ02
e= , with r0 = , (21)
GM ψ2 GM ψ2

r0 in [cm] is the radius of the circular orbit. It behaves like 1/λ, thus increasing with time.
For elliptical solutions, with e < 1, the semi-major and semi-minor axes, respectively, are
a = (1−r0e2 ) and b = √ r0 2 . These are not classical conics with constant parameters (except
1− e
for e). As the result of the time dependence of r0 , they are expanding conics [24]. From (21),
ṙ Ṁ
one has, considering a distance r which may apply to a conic or a circle, r = −M − 2 ψ̇ψ .
Since M ∼ t, and given Equation 4, one has

τ Ṁ ψ ṙ ψ
M(τ ) = M (τ0 ) [tin + (1 − tin )], = , and = , (22)
τ0 M τ0 r τ0

with ψ̇ = −ψ2 /τ0 . This implies that the distance r, or the semi-major axis a(τ ) of the
elliptical motions, behaves like r (τ ) = r (τ0 ) [tin + ττ0 (1 − tin )]. The semi-major axis is
therefore increasing linearly in time τ, consistent with the variation of a(t) = a(t0 ) × t
obtained in the t-scale [24].
We now turn to the orbital period T, which, for simplicity, is considered in the circular
case. With the angular velocity ϑ̇ = 2π/T (for an elliptical orbit ϑ̇ would be the mean
angular velocity) and the angular momentum conservation ϑ̇ = L τ0
r2 ψ
, one is led to

ϑ̇ ṙ ψ̇ ψ ψ ψ Ṫ ϑ̇ ψ
= −2 − = −2 + =− , thus =− = . (23)
ϑ r ψ τ0 τ0 τ0 T ϑ τ0

This implies that the period T is varying, like T (τ ) = T (τ0 ) [tin + ττ0 (1 − tin )]. Thus,
the mass M, the parameters a, b, and r0 , and the orbital period T all have the same kind of
time dependence.
Symmetry 2024, 1, 0 9 of 25

Let us now turn to the orbital velocity. Starting from (15), for the orbital velocity, we
have υ = r ϑ̇ = Lψτr0 . Again considering a circular orbit, by introducing the expression of r0
given in Equation (20), one obtains

L τ0 GMψ2 GMψ G 2 M 2 ψ2 GM
υ = r0 ϑ̇ = 2
= , and υ2 = 2
= . (24)
2
ψ L τ0 Lτ0 2
L τ0 r0

The time dependencies of M and ψ cancel each other out; we have used (20). The
orbital velocity has the usual standard expression; it is scale-invariant, keeping constant in
time during secular orbital expansion. This is consistent with the fact that the potential is a
conserved quantity in the theory. However, it does not imply that the velocity dispersion is
the same, as seen in Section 3. Regarding Equation (24), it is interesting to note that there
is in fact a very subtle interplay in the scale-invariant theory: the tangential (additional)
ψ dr
dynamical acceleration τ00 dτ exactly compensates for the usual slowing down due to the
orbital expansion so as to keep the orbital velocity constant during the secular expansion.

4.2. Recent Observations


The bulk of observations of spiral galaxies show flat rotation curves up to galactic
distances of a few tens of kpc according to Sofue and Rubin [43]. The results by Huang et
al. [44] show an extended rotation curve up to about 100 kpc based on the observations of
about 16,000 red clump giants in the outer disk and 5700 K giants in the halo (Figure 2). A
proper account of the anisotropic motions is performed in the analysis, which also permits
one to see the undulations due to spiral arms in the inner region; but, on the whole, this
part could be regarded as globally flat. Beyond 26 kpc a progressive decrease is observed
up to distances as large as 100 kpc.
More recent results by Eilers et al. [45] and by Jiao et al. [46] find steeper curves for
the 5 to 25 kpc inner distance interval. The study by Eilers is based on the 6D data (location
and velocities) of about 23,000 red giant stars where the velocity curve is determined from
the Jeans equation for an axisymmetric gravitational potential. These authors found a
decreasing velocity of 1.7 (±0.1) km/s per kpc. They estimated a value of 7.25 (±0.26) ×
1011 M⊙ for the total mass of the Milky Way, and found that dark matter is dominating
for radii larger than about 14 kpc. The results by [46] are based on Gaia DR3. They find a
steeper decrease in the velocity of 30 km s−1 between 19.5 and 26.5 kpc, leading to a much
(+0.24)
smaller mass value of 2.06(−0.13) × 1011 M⊙ . Up to 20 kpc from the center, the data by both
groups agree with the lowest points of the wiggles due to galactic arms by Huang et al.,
while the wiggles are absent from the other two axisymmetric studies (Figure 2).

4.3. Predicted Effects in the Galactic Evolution


The above secular variations and velocity constancy allow us to reconstruct the evolu-
tion of galaxies, for the part due to scale invariance [25,47]. In particular, we consider the
two following effects:
• The orbital radius (here the galactocentric radius) increases with time (Equation (22)).
• The orbital velocity υ of a given star around the galactic center remains constant
during the additional expansion (Equation (24)).
The two above effects contribute to shape the rotation curves of galaxies starting
from the initial velocity profile υ(r ). From the initial profile, likely not far from a steep
Keplerian shape, the above effects are progressively producing a flattening of the curve,
which increases with the age of the galaxy and also depends on the other internal dynamical
effects. A clear prediction is that the steepness of the profile should increase with redshifts z.
The time intervals considered in the evolution of galaxies are not negligible with
respect to the age of the Universe; thus, we cannot simply write

∆r r (τ )
= ψ0 0 , (25)
∆τ τ0
Symmetry 2024, 1, 0 10 of 25

as would be performed for small changes at the present time. Instead, we must integrate
Equation (22) over time τ between τ1 (in Gyr) the epoch considered in the past with size r1 ,
and the present time τ0 with size rτ0 ,

(1 − Ωm )1/3
Z r0
dr dτ dτ dτ
Z τ0 Z τ0 Z τ0
= = 1/3
= , (26)
r1 r τ0 τ1 Ω1/3
m + τ 1/3
τ0 (1 − Ωm )
τ1 τ0 (1−ΩΩm )1/3 +τ τ1 τΩ + τ
m

where we use the constant τΩ defined in Equation (9). The integration gives

r0 τ + τ0
= Ω . (27)
r1 τΩ + τ1

For example, for Ωm = 0.045 [23], τ0 = 13.8 Gyr and an age τ1 of 6 Gyr, one has
τΩ = 7.62 Gyr and a scaling of the distances rr0 = 1.57. This means that, at an age of 6
1
Gyr, the same velocity as one on the red curve of (Figure 2) was reached at a galactocentric
distance r1 , which is a factor 1.57 smaller than the present one. As an illustration of the
predicted effects, we take the rotation curve given by Huang (2016) [44] (red curve in Figure
2). The above two rules lead to the construction of the past rotation curves of galaxies at
different past ages of 0.2, 3, 6, and 9 Gyr, as shown by the blue curves in this figure. We
notice that they become steeper for further in the past. Among the blue curves, the left one
corresponds to the early age of galaxy formation (0.2 Gyr). When more detailed curves may
come, the above process may also be applied and used for comparisons with the curves in
high-redshift galaxies (see below where some observable trends are already noticed). These
results illustrate the cosmic evolution resulting from scale-invariant effects.

260 Huang et al. 2016 ( mean )


Vc Eilers et al. 2019
Jiao et al. 2023
(km s-1)
Rotation curves in the past (SIV)
220
Rotation curve in r - 0.5

180 Present
13.8 Gyr

140
Ages: 0.2 Gyr 3 Gyr 6 Gyr 9 Gyr
R(kpc)
100
0 10 20 30 40 50 60 70 80 90 100

Figure 2. Evolution of the rotation curve of the Milky Way. The gray points are the velocity averages
observed by Huang et al. [44], with their error bars. The thick red line represents a mean rotation
curve. The thin dashed red line describes the flat mean of the wiggles of the velocity distribution up
to a radius of 26 kpc. The brown and orange lines show, respectively, the results of Eilers et al. [45]
and by Jiao et al. [46] for the inner galaxy. The blue lines show the rotation curves predicted by the
scale-invariant theory for different ages in the past history of the Universe, starting backwards from
the red curve. Calculations have been performed with no dark matter in a model with Ωm = Ωb ,

where Ωb = 0.045. The thick dashed green line shows a Keplerian curve in 1/ r near the age of
galaxy formation.

The rather steep—but short—curves in the inner part, found more recently by Eilers
et al. [45] and by Jiao et al. [46], could also be treated backwards in the same way as for
Huang data. They would lead to a very small core with a steep velocity gradient in the
early evolutionary stages. We also note that these recent curves rather correspond to the
minimum velocity of the dips associated with spiral arms. At high redshift, we may expect
the following effects of scale invariance in galaxies:
Symmetry 2024, 1, 0 11 of 25

1. In the past, the extent of galaxies was smaller than today. Galaxies were more compact
according to the ratio given by Equation (27).
2. The cosmic expansion associated with scale invariance flattens the rotation curves as
galaxies age.
3. In the past, the rotation curves should have steeper decrease and closer to the galactic
center than today, thus suggesting smaller fractions of dark matter in the past. This of
course does not preclude other current effects of stellar dynamics to operate over the
ages.

4.4. Some Recent Results in Redshifted Galaxies


It is interesting to compare these theoretical results with some recent observations. Six
star-forming galaxies with redshifts between z = 0.8 and 2.4 were analyzed by Genzel et
al. [48], and a large sample of 101 other galaxies from the KMOS and from the SINS/zC-
SINF surveys with redshifts between z = 0.6 and 2.6 were studied by Lang et al. [49].
Both surveys give trends corresponding to the above point #3: the observed past rotation
curves were not flat, but were displaying steeply decreasing velocity curves. Thus, even
at distances of several effective radii, the authors found that the dark matter fractions
were modest in the past. A variety of interpretations of the above results have taken place
(see also [50]). However, the further analysis by Genzel et al. [50] from the same survey
confirmed that at z ∼ 2 galaxies have very low dark matter fractions, below 20%.
The observations are also improving rapidly, and very recently the above results
were confirmed by new data from different instruments and wavelengths. In particular,
a hundred galaxies at the peak of their star formation (z = 0.6 − 2.5) were observed with
ESO-VLT, LUCI, NOEMA at IRAM, and ALMA, and these were further analyzed, doubling
the previous samples [51]. These new observations show rather steep rotation curves,
which—according to the authors—demonstrate that the fractions of dark matter strongly
decrease with increasing redshifts: for example, at z = 0.85, the fraction of dark matter is
estimated to be 38% of the typical value, and this fraction goes down to 17% at z = 2.44.
In standard models with dark matter, the above decrease of DM with redshifts is difficult
to interpret. We do not know where, at high z, the dark matter is located, in order to enable
such small dynamical signatures only. The possibility that the missing dark matter is lying
in more external regions would be a rather troublesome one. Indeed, in the CDM theory
the baryons set in the potential wells already created by dark matter during the radiative
era. Thus, it is not clear why galaxies would contain less dark matter in the past, since the
accumulation of DM is considered preceding that of baryons. Another difficulty is that,
contrarily to the above results showing less dark matter at higher z, some results on lensing,
in particular for JWST-ER1, seem to suggest more dark matter for high-redshift galaxies, as
will be seen in Section 6. The two above properties represent some internal contradiction for
the dark matter interpretation, and this tends to support our conclusion that the dark matter
hypothesis is likely not the right solution for all problems in the dynamics of galaxies.

4.5. The Connection between Dark Matter and Baryons: The Signature of a Gravity Effect
There is an interesting observation about the radial acceleration (RAR) studied by
2
Lelli et al. [52], who showed that the relation between the average accelerations υR and GMR2
for galaxies deviates from a 1:1 relation for gravities below 1.2 × 10−8 cm s−2 . The RAR is
accounted for by both MOND [52] and by the SIV theory [53]. As found by [52] the amount
of dark matter (DM, measured by the deviations from the 1:1 line) seems to be always
directly determined by the amount of baryonic matter, although 2700 measurements have
been made in different galaxies and at different distances from their centers. Such a result is
pointing in favor of a gravity effect rather than for DM.
The same kind of relation between the dynamical and baryonic masses of central
regions is also present in clusters of galaxies according to Chan (2022) [41]. Moreover, such
a relation is also observed in redshifted galaxies, where Nestor Shachar et al. [51] find
a strong correlation between the fraction of DM and the baryon surface density within
Symmetry 2024, 1, 0 12 of 25

the effective radius (i.e., the part corresponding to the flat curve). The convergence of
such findings in three very different environments confirms the general intimate relation
between baryons and the assumed dark matter. It is clearly supporting the views by Lelli
et al. that this general connection is the signature of gravity effect.
To summarize, the higher velocities in galaxy clusters, the flatter rotation curves at
low redshifts than at high redshifts; the omnipresent connections between the distribution
of baryons and of the supposed dark matter are all in support of a dynamical evolution at
low densities governed by scale-invariant effects in the dynamics of galaxies.

5. Wide Binary Stars


The effects of scale invariance should also be active in low-gravity systems with long
timescales. An interesting case is that of very wide binary stars (wide binaries), where new
interesting results from the astrometric Gaia satellite have recently come out. As is often
the case for physical effects at the limit of experimental possibilities, this domain is rich in
different and controversial views.

5.1. Current Status of Observational Results on Very Wide Binaries


Among recent results, we mention the study by Hernandez [54], who studied a
sample of 929 carefully selected binary pairs from the Gaia Early Data Release 3 eDR3
[55] with high quality requirements. He found that, below a0 ≃ 1.2 × 10−10 m s−2 , the
velocities remain constant, independently of the orbital radius. This value a0 corresponds

to the deep MOND limit, below which gravity behaves like g = gN a0 (with gN the
Newtonian gravity), and orbital velocities remain constant according to [56,57]. Another
study was published by Hernandez [58] on the basis of the DR3 Gaia catalog with 800,000
binary candidates. Hernandez considered the binary probability B p for each star based on
spectral, photometric and astrometric information, the RUWE index, the radial velocities,
the isolation criteria, the Gaia FLAME work package for mass estimate, etc. These very
stringent quality criteria gleaned a set of 450 top-quality isolated binaries, which do well to
confirm the earlier results.
Pittordis and Sutherland [59] studied a large sample of 73,159 wide binaries from the
Gaia eDR3 objects within 300 pc and with a G magnitude brighter than 17. With masses
determined by a mass–luminosity relation, they analyzed the distributions of the relative
projected velocities and found a peak close to the Newtonian predictions, in addition to a
long tail, likely produced by unseen objects. They noted, with some reserve, that standard
gravity is preferred over MOND.
Chae [60] developed a Monte Carlo method to de-project the observed 2D data. He
proceeded to a calibration of the fraction f multi of multiples for gravities > 10−10 m s−2
and found a fraction of 0.3 to 0.5 in agreement with current results.The comparison of the
kinematic acceleration g = υ2 /r and the Newtonian gravity shows a ratio g/gN = 1.43 at
gN = 10−10.15 m s−2 in agreement with one of the MOND generalizations. Some critical
remarks on other works are mentioned.
Banik, Pittordis, and Sutherland et al. [61] performed tests of the MOND dynamics
with wide binaries, from a highly selected sample of 8611 wide binaries from Gaia DR3
within 250 pc of the Sun. They proceeded to a modeling of the orbits in the Newton
and MOND approximations. Their statistical analysis favored a strong preference for
Newtonian gravity, although their best model does not fully reproduce the observations.
They also suggested further modifications of MOND.
A new study by Chae [62] uses the catalog by El-Badry et al. [63] that provides a
probability, ℜ, of chance alignment. He applied, among others, the severe requirement
that ℜ < 0.01, i.e., mandating “pure binaries”. He pointed out that the estimate f multi by
[61] of f multi = 0.70 was incompatible with the range currently obtained. He studied the
distributions of the sky-projected velocities as a function of the sky-projected separation s
for pure binaries. Deviations from Newton’s starts around log s = 3.3 (in au), increasing
Symmetry 2024, 1, 0 13 of 25

up to log s = 3.8, keeping about constant deviation by about 0.1 dex larger (Figure 3), a
property which appears to differ from MOND’s.

5.2. Dynamical Evolution of Loose and Detached Systems


For increasing distances from a star of mass M, the attraction of the other star in a
binary or of the galactic field makes gravity vanish. This occurs at the Jacobi radius, or
Roche radius, where the galactic field becomes dominant; a value of ∼ 1.70 pc is estimated
ψ
for solar mass stars [64]. Before this limit, the dynamical acceleration τ0 × υ may overcome
the Newtonian gravity gN . The equality of the two terms occurs at [17]:

(1 − Ωm )2 n c (1 − Ωm )(1 − Ω1/3
m )
g0 = n c H0 or g0 = . (28)
4 2 τ0

The product c × H0 is equal to 6.80 × 10−8 cm s−2 . For Ωm =0, 0.10, 0.20, and 0.30, we
obtain g0 = (1.70, 1.36, 1.09, 0.83) × 10−8 cm s−2 , respectively. These values of g0 cover the
domain of a0 . For gravities below g0 , the equation of motion (5) accepts, as an approximation,

a relation of the form g = gN g0 [56], as discussed above; it is valid over the last 400 Myr
to an accuracy of 1-2% depending on Ωm [17]. For a given total mass Mtot of a binary, the
equality of the two terms occurs at a separation:
s
G Mtot
strans = , for Mtot = 1.0 M⊙ , strans = 7.0 [kau] or 0.034 [pc]. (29)
g0

What happens for wide binaries near the transition? For a decrease in effective gravity,
the eccentricity e and radius r0 of the circular orbit are tending to ∞ from Equations (20)
and (21). The orbit extends with an increase in the ellipticity up to e = 1, where it becomes
parabolic. At this limit, we have

GM ψ2 1
C = ≡ . (30)
L2 τ02 r0

For a further decline in gravity, e becomes larger than 1 and the orbit is hyperbolic,
tending to a straight line. Let us examine the behavior of the orbital velocity at this limit:

d2 r ψ(τ ) dr dυ ψ(τ ) ψ̇ dψ
2
= , or = dτ = − dτ = − , with ψ̇ = −ψ2 /τ0 .
dτ τ0 dτ υ τ0 ψ ψ
(31)

k k τ
Integration gives ln υ = − ln ψ + const. thus υ = = [t + (1 − tin )] ,
ψ (1 − tin ) in τ0
(32)
where k is a constant. Setting τ = τ0 , we have υ(τ0 ) = (1−kt ) and υ(τ ) becomes
in
υ(τ ) = υ(τ0 ) [tin + ττ0 (1 − tin )]. The velocity no longer remains constant with secular vari-
ations as in the bound system, but it increases linearly in time τ under the effect of the
ψ(τ )
dynamical acceleration τ0 υ. Turning to the distance, r, covered during this motion, we have
from (32),

dr τ 1 τ0 τ
= υ(τ0 ) [tin + (1 − tin )], thus r (τ ) = υ(τ0 ) [tin + (1 − tin )]2 . (33)
dτ τ0 2 (1 − tin ) τ0
ψ
Thus, when the dynamical gravity τ0 υ dominates, the motion becomes hyperbolic,
with e > 1, a velocity increasing linearly with time and a covered distance with a quadratic
dependence in time, r (τ ) = r (τ0 )[tin + ττ0 (1 − tin )]2 . We call this regime the dynamical
regime. In order to numerically estimate the changes in velocities and covered distances, we
Symmetry 2024, 1, 0 14 of 25

consider a reference time τN marking the transition of the Newtonian regime and to the
dynamical one. We then obtain the following ratios:

υ(τ0 ) 1 r (τ0 ) 1
= τN , and = τN 2
. (34)
τ0 (1 − tin )] τ0 (1 − tin )]
υ(τN ) [tin + r (τN ) [tin +

This gives the change in velocity and radius, between the transition time τN and today
in the dynamical regime, for binary systems in the Milky Way.

5.3. Comparison with the Analysis by Chae [62]


log 10 p (km s-1)

5 Gyr
3 Gyr
1 Gyr
Effects of the
shifts in  and r

Original from Chae (2024)

Figure 3. Projected velocities υ p as a function of separation s for the main sample in Figure 13 from
Chae [62]. The very small red dots are the observed values and the blue dots are the Newtonian
values in the Monte-Carlo simulations. The larger dots, red (obs.) and blue (simul.) and their
connecting lines are the medians and percentiles, as indicated. The central black dashed line shows

the Keplerian relation in 1/ s. The green lines are additional indications of the mean deviations
along the hyperbolic path of the loose systems after 1, 3, and 5 Gyr from the time, τN , the transition
from Newtonian to dynamical acceleration. The direction of the effects in velocity and separation is
indicated by a green arrow at the bottom left. The deviation of a given system increases linearly with
time as indicated by Equation (34). The mean observed relation for the loose systems corresponds to
an evolution during about 2 to 3 Gyr. In a few Gyrs, loose systems drift away from the Newtonian
relation in a way compatible with the dynamical evolution. (Figure adapted from Chae [62]).

Figure 3 shows in red the relation between log υ and log s for a sample of 2463 binary
pairs with masses derived from the absolute G-band magnitude, MG , see comments. From
an observed set of M and s values, each Monte-Carlo (MC) realization draws the velocity
from the classical Newtonian equations with the appropriate statistical distributions of the
orbital parameters. Series of 400 MC simulations are performed to estimate statistical error
bars (blue line). The average total mass of the binary systems is around 1.5 M⊙ ; thus, the
main sequence lifetimes are of several Gyr. We consider systems starting deviating from
Symmetry 2024, 1, 0 15 of 25

Newton’s Law at 1, 3 ,and 5 Gyr ago, i.e., at past ages τN = 12.8, 10.8, and 8.8 Gyr. Thus,
from Equation (34), the shifts in velocities and distances since time τN , respectively, are:
   
υ(τ0 ) r (τ0 )
log = 0.020, 0.064, 0.113, and log = 0.040, 0.128, 0.226. (35)
υ(τN ) r (τN )

In Figure 3, these shifts produce displacements from √ the red to the green lines,
for binaries at different locations on the reference line υ ∼ 1/ s. The mean shift increases
between s = 3 [kau] and 8 [kau], in agreement with strans . The observed shifts correspond
to an evolution in the free regime for about 2 to 3 Gyr. These shifts are clearly larger than
the error bars. We note that all the possible variations/combinations of the selection criteria
in the sampling by [62] give rather similar shifts.
In Figure 4, Chae [62] analyzed a smaller sample including only the 40 binaries with
uncertainties < 0.5 %, comparable to astrometry. Three bins are formed; the highest s-bin
shows a large deviation from Newtonian relation, compatible with free motion during
about 5 Gyr. This extreme sample supports the existence of significant deviations from
Newtonian relation for log s ≥ 3.5 and above. It appears as a robust conclusion that
the direction and size of the observed effects are compatible with a dynamical evolution
dominated by the additional scale invariance term for several Gyr.

log u
[km s-1] Newtonian relation u ~ s-1/2
Obs. spatial velocities
Newtonian MC realization
0.2

0.0

-0.2

Isochrones
-0.4
5 Gyr
3 Gyr
-0.6 Data from Chae (2024) 1 Gyr

-0.8
2.5 3.0 3.5 4.0 log s [au]

Figure 4. Comparisons of observations and theory for a sample of 40 very wide binaries with
exceptionally precise radial velocities, with individual relative errors smaller than 0.005, as selected
by Chae [62]. The green broken lines shows the isochrones corresponding to departures from the
Newtonian law after 1, 3, and 5 Gyr of evolution under the dynamical acceleration in the scale-
invariant theory. The departure from Newton’s relation is progressive and tends towards a value
between 3 and 5 Gyr (data are from Chae [62]).

5.4. Comparison with the Analysis by Hernandez [58]


The sample of 450 wide binaries by Hernandez [58] results from stringent quality
criteria (Section 5.1). Mass estimates make use of an internal Gaia FLAME work-package
based spectroscopic information in DR3. The relative velocity ∆V analysis in the plane of
the sky was primarily determined through Gaia proper motions in two perpendicular directions.
The binned root-mean-square ⟨∆V ⟩ of the relative velocities between the two components
Symmetry 2024, 1, 0 16 of 25

of selected pairs are shown in Figure 5 as a function of the 2D projected separations s


between the two components of each binary pair.
Figure 5 shows the results. The blue line shows the Newtonian relation υ ∼ s−1/2
from simulations of 50,000 binary pairs of two 1 M⊙ mass [64]. This line follows a Keplerian
relation ⟨∆V ⟩ ∼ s−1/2 up to the Jacobi radius, well beyond the value of about 0.01 pc (note
that 0.01 pc corresponds to 2.063 kau).
The Keplerian relation is very well followed for separations up to about 10−2 pc.
However, for larger separations, the mean relative velocities significantly deviate from the
Keplerian relation and tend toward an essentially flat constant value (see Figure 5).
The comparison of the results by Hernandez [58] with isochrones showing the
duration of the free evolution also supports an evolution in the dynamical regime of the
order 5 Gyr, in agreement with the best-quality results in Figure 4. The key point is that
these different results on very wide binaries are supporting significant deviations from the
standard Newtonian predictions, consistent with an evolution dominated by Equation (33)
for several Gyr.
Relative velocity in plane of sky [km/s]

1.0 136 Isochrones 5 Gyr


111 135 3 Gyr
1 Gyr
114 81
64 58 43 24
78
41

Data from Hernandez (2023)


0.1
10-3 10-2
2D-separation s [pc]

Figure 5. Binned root mean square ⟨∆V ⟩ of the relative velocities in the plane of sky as a function of
the 2D projected separation s for 450 binary pairs by Hernandez [58]. There is a partial overlap of
the binned pairs. The number of binary pairs in the various means are indicated. The green lines
show the isochrones corresponding to departures from the Newtonian law after 1, 3, and 5 Gyr of
evolution (adapted from [58]).

A possible difficulty has been raised by Loeb [65]: in wide binaries, the gravitational
redshift has about the same amplitude as the Doppler–Fizeau effect. This might be a problem;
however, it is not necessarily a problem if we consider the differential effects. Following
Loeb, we express the velocity shift by gravitation υGR = − GcRM ; with the mass–radius relation
 0.8  0.2
R M M
R⊙ ≈ M⊙ for solar-mass stars, it becomes υGR ≈ − 0.6 M⊙ . For a mean a mean
mass ratio M2 /M1 of about 0.5 to power of 0.2, this gives a value of 0.87 for the ratio of the
gravitational redshifts of the two components, i.e., typically a velocity difference of about 0.08
km/s which is small for a mean velocity of 0.60 km/s. Thus, the differences of gravitational
redshifts between the two components are much smaller than the velocity shifts. Furthermore,
in the data by Hernandez, the relative motions are mainly derived from the proper motions
in the plane of the sky, where gravitational effects are not in action. Moreover, gravitational
redshifts do not introduce any systematic bias in the mean of many binaries.
As a conclusion, the recent results from Gaia DR3 for very wide binaries by [58]
and [62] show orbital velocities larger than Newtonian for separations larger than about 3
kau, but they are in agreement with the predictions of the scale-invariant theory. It seems
Symmetry 2024, 1, 0 17 of 25

difficult to advocate for a halo of dark matter surrounding each binary system, as currently
performed in the domain of galaxies and clusters.

6. The Discrepancy between the Lensing and the Spectroscopic Mass of Galaxies
Gravitational lensing is a powerful tool of astrophysics. It is no exception to
the current problems related to dark matter, in the sense that the masses from lensing
are generally larger those from photometry, see Sections 6.3 and 6.4. Thus, it is worth
examining the effects of scale invariance in the context of lensing.

6.1. The Geodesic of Deflected Light Rays in the Scale-Invariant Theory


To calculate the modified trajectories of a body near another mass, the starting
point is the equation of the geodesic [2,15]:

duα
+ Γαµν uµ uν − κµ uµ uα + κ α = 0. (36)
ds
For the metric of the system, we adopt the Kottler metric [66]:

−dr2 2GM Λr2 2 2


 
ds2 = − r2 (dϑ2 + sin2 ϑdφ2 ) + 1 − − c dt . (37)
2GM Λr2 rc2 3
1− rc2
− 3

Λ = λ2 ΛE , withΛE the cosmological constant in GR; λ is the scale factor, a slowly


variable universal function. We calculate the geodesic for a body of mass m = GM/c2 ; the
case of light rays is specified later. The geodesic differs from the classical case of GR by the
terms −κµ uµ uα + κ α :

dr dt dr
α=1 : −κ µ u µ → −κ0 , (38)
ds ds ds
dθ dt dθ
α=2 : −κ µ u µ → −κ0 ,
ds ds ds
dϕ dt dϕ
α=3 : −κ µ u µ → −κ0 ,
ds ds ds
dt dt dt κ
α=0 : −κ µ u µ + κ0 → −κ0 + 20 .
ds ds ds c

With κ0 = −λ̇/λ = 1/t. We adopt a simplified form of the metric:

ds2 = −eλ dr2 − r2 (dθ 2 + sin2 θdϕ2 ) + eν c2 dt2 ;


2m Λr2
 
1
with eλ =
Λr2
, e ν
= 1 − − , (39)
1 − 2m − r 3
r 3

where m = GM/(c2 ) has the units of a length and is subject to slow time-variations
according to Equation (22). We also assume, as usual, that the motions occur in a plane
with θ = π/2, and we obtain

2 2 2
d2 r 1 dλ eν−λ dν c2 dt
  
dr dϕ dt dr
+ − re−λ + − κ0 = 0,
ds2 2 dr ds ds 2 dr ds ds ds
 2
d2 ϕ 2 dr dϕ dt dϕ d2 t dν dt dt κ
2
+ − κ0 = 0, and 2
+ − κ0 + 20 = 0, (40)
ds r ds ds ds ds ds ds ds ds c
Symmetry 2024, 1, 0 18 of 25


The last two equations may be easily integrated. Let us call υ = ds ; for the first
one, we obtain

dυ 2 dr 1 dt dυ dr dt dϕ
+ υ− υ = 0, and +2 − = 0, giving r2 = ht. (41)
ds r ds t ds υ r t ds
Here, h is an integration constant, representing the angular momentum (divided by c) per
mass unit. This is the modified conservation law, which increases with time t/t0 . This
is consistent with the scale-invariant two-body problem [24,25], where the orbital radius

slowly increases, while the orbital velocity r ds keeps constant, see Section 4.1 . Turning to
 2
the third equation of (40), we multiply it by ds dt and obtain

2
d2 t dν dt

κ ds
+ − κ0 + 20 = 0. (42)
dt2 dt dt c dt

The first two terms are zero, the second one because ν is scale-invariant, independent of
time, since in the parenthesis of (37), both mass and radius are scaling like 1/λ, so that
2m/r is scale-invariant; the same is the case for Λ, scaling like λ2 . The two terms with κ0
are also equal to zero in the t-equation. Thus, one is just brought back to the classical case;
calling upon w = ds/dt, one is left with

dw dν
+ w = 0, with solution w = dt/ds = k e−ν(s) , k = const. (43)
ds ds
Now, we turn to the more difficult Equation (40). Let us quote [11], who said
about the corresponding equation in the GR context, “Instead of troubling to integrate (40) we
can use in place of it (39)” (the labels of both equations have been changed appropriately).
Thus, from (39) we obtain
 2  2  2
dr dϕ dt
eλ + r2 − e ν c2 + 1 = 0. (44)
ds ds ds

With the above expression for w, the third term can also be written −eν c2 k2 e−2ν = −c2 k2 e−ν =
c2 k 2
− 2GM Λr2
. Thus, Equation (44) becomes
[1− − ]
rc2 3

2 2
Λr2
     
dr dϕ 2m dϕ dϕ
+ r2 − 1 + r 2 ( )2 − 1 + r2 ( )2 = c2 k2 − 1. (45)
ds ds r ds 3 ds

Eliminating ( ds ) with the last expression of (41), we obtain
2
h2 t2 2m 2mh2 t2 Λr2 Λh2 t2

ht dr 2 2
+ = c k − 1 + + + + . (46)
r2 dϕ r2 r r3 3 3

We divide it by h2 t2 and set u = 1/r for consistency with the classical Binet equation given
by Equation (19):
2
c2 k2 − 1 2mu Λ Λ

du
+ u2 = 2 2
+ 2 2 + 2mu3 + 2 2 2 + . (47)
dϕ h t h t 3u h t 3
Now, we derive this equation with respect to ϕ and obtain

d2 u c2 k 2 − 1 d
     
du du 2m du 12u dm 2mu d 1
2 + 2u = + 22 ++ 2 2 + 2 +
dϕ dϕ2 dϕ h2 dϕ h t dϕ t2
h t dϕ h dϕ t2
Λ λ2 1 du Λ d λ2
 
du dm 2ΛE λ dλ
6mu2 + 2u3 − 2 E2 2 3 + 2E 2 2
+ . (48)
dϕ dϕ 3h t u dϕ 3h u dϕ t 3 dϕ
Symmetry 2024, 1, 0 19 of 25

There is a direct relation between ϕ and time since the time t is uniquely defined
along the geodesic described by the variation of angle ϕ. Thus, λ, which is a function of

time t, is also a function of ϕ. Let us now multiply the result by 12 du :

d2 u c2 k 2 − 1 d
   
m1 u dm mu d 1
+u = + + +
dϕ2 2h2 du t2
h2 t2 h2 t2 du h2 du t2
dm ΛE λ2 1 Λ d λ2 Λ λ dλ
 
+ 3mu2 + u3 − 2 2 3 + 2E 2 + E , (49)
du 3h t u 6h u du t2 3 du

This equation combines the effects of general relativity and scale invariance. It
can be used to study the advance of perihelion and other post-Newtonian effects in the
scale-invariant context.

6.2. The Invariance of Lensing to Scale Transformations


For a light ray, we have the additional condition ds = 0, i.e., its proper time never
changes. According to Equation (41), this implies h = ∞, see [11]. This simplifies Equation
(49), where only three terms are remaining on the right side. From the gauging condition
2
3 λ̇λ2 = c2 λ2 ΛE [1,4] (with c2 for the consistency of the units), we obtain, with λ = t0 /t in
the units of the cosmological models, ΛE = c23t2 ; in current units, ΛE = c23τ2 with τ0 the
0 0
age of the Universe (see also [21] for a discussion of these properties). Thus, Equation (49)
becomes !
d2 u dm 1 dλ
+ u = 3mu2 + u3 + λ . (50)
dϕ2 du c2 τ02 du

The scale factor λ is subject to the current limitations (Section 2). If λ is a constant,
we again have the usual equation for lensing in GR. cτ0 is a distance of the order of the
radius of the Hubble sphere c/H0 . Numerical integration of the scale-invariant geodesics,
for the specific case of the z L ∼
= 2.0 lens galaxy in the extreme system JWST-ER1 (Section 6.3)
shows that the two additional terms are negligible, with about O (10−6 ) of the Einstein term.
Thus, we may conclude that the photon’s geodesics remain unchanged and the deflection
angle α = 4GM/(c2 R) is not affected by scale-invariant effects, noting also that the mass
M and radius R vary the same way with time.

6.3. The Problem of the Photometric Masses: The Case of JWST-ER1


We first consider the case of JWST-ER1, a strongly lensed galaxy with a complete
Einstein ring and a high lensing-to-stellar mass ratio [67]. The lens redshift is zL = 1.94; it
is a massive quiescent galaxy with a low star-formation rate and an age of τL = 1.9+ 0.3
−0.6 Gyr,
implying mainly low-mass stars. Its unit-less age is tL = 0.62, according to
h i1/3
t = Ωm + (1 + z)−3/2 (1 − Ωm ) . (51)

as derived from (2) with Ωm = 0.05, i.e., assuming no dark matter. According to the
scaling, M ∼ t, this implies that masses in JWST-ER1 are a fraction 0.62 of the present
ones. This shift in mass is rather similar to the solution proposed by van Dokkum et
al. [67] to solve the discrepancy between the stellar mass (M⋆ = 1.1 × 1011 M⊙ ) and the
lensing mass (MTot = 6.5+ 3.7
−1.5 × 10
11 M ). They suggested a shift from Chabrier’s IMF

[68], with few low-mass stars giving M∗ = 1.1+ 0.2 11
−0.3 × 10 M⊙ , to a very bottom-heavy IMF
+0.6 11
(“Super-Salpeter IMF”), giving M∗ = 4.0−0.8 × 10 M⊙ . The standard Salpeter IMF gives
an intermediate value of M∗ = 2.0+ 0.5 11
−0.5 × 10 M⊙ . Figure 6 illustrates the various IMF’s.
New results on JWST-ER1 by Mercier et al. [69] give a much higher redshift
zS = 5.48 ± 0.06 of the source instead of 2.98 by [67], while zL = 2.02 ± 0.02 is about
the same. The estimated masses are also different; they are, respectively, MTot = (3.66 ±
0.36) × 1011 M⊙ and a stellar mass of M∗ = 1.37+ 0.14 11
−0.11 × 10 M⊙ , which partially reduces
Symmetry 2024, 1, 0 20 of 25

the discrepancy. Even with the Super-Salpeter assumption, some fraction of dark matter is
necessary, for both studies [67,69].
Since the deflection angle is free from scale-invariant effects, we now turn to the
mass estimates from photometry. The observed luminosity Lobs reflects the stellar masses at
the age we see them. For an age τL = 1.9+ 0.3
−0.6 Gyr assigned by Prospector (PROSPECTOR is a
flexible code for inferring stellar population parameters from photometry and spectroscopy
spanning UV through IR wavelengths [70]), the turnoff mass is around 1.7 M⊙ [71], in
agreement with a remark by [72] that most of the light of the old populations in galaxies
comes from stellar masses in the range of [1 − 2] M⊙ . However, a shift in mass by a factor
0.62 for JWST-ER1 implies a big change in the relative number frequency of stars with
respect to the standard case and has great incidence.
Let us estimate the effect of such a shift on the average mass–luminosity ratio
⟨ M/L⟩ of a galaxy. As an example, we take an initial mass function ξ (ln m) of the Salpeter
form [73]:
dN dN dN
ξ (ln m) = = m = Am− x , or = Am−(1+ x) , (52)
d ln m dm dm
with x = 1.35 (see Figure 6). The mass of the galaxy then writes

dN A
Z Z
Mgal = m dm = A m− x dm = | ( m − x +1 ) | m
mmin ≃ Amm ,
max (53)
dm −x + 1
where A = const. and mm the mean mass over the IMF (mm =13.7 M⊙ for a mass range
from 0.01 to 100 M⊙ ). If we call ℓ(m) the luminosity of a star of mass m, the stellar mass–
luminosity relation writes ℓ(m) = B mα , where B is another constant. The luminosity of a
galaxy is

dN AB α− x mmax
Z Z
L gal = ℓ(m) dm = AB mα m−( x+1) dm = | m |mmin = AB m L , (54)
dm α−x
where m L is the mean mass over the luminosity distribution. Typical values are 1.7–1.8 M⊙ .
Thus, the mean mass–luminosity ratio of a galaxy scales like

( α − x ) 1− x x − α
 
M
= m m . (55)
L B (1 − x )

where m is a representative mass; the exact value is not critical, since all masses are reduced
by the same factor. The exponents α are large, hence making ℓ a steep function of m (Figure
6). We have ℓ ∼ mα ,

α = 3.69, m ∈ [9 − 2] M⊙ α = 4.56, m ∈ [2 − 1] M⊙ α = 4.00, m ∈ [1 − 0.6] M⊙ ,


(56)
according to [71]. The individual m/ℓ ratios are varying with power (1 − α) of the mass
m, while Equation (55) shows that the factors are different when integrated over the mass
spectrum of a galaxy. Now, we can see what is doing a change in the mass by a factor
f = 0.62 (or 0.614 by [69]). The ratio of the modified overall mass–luminosity ratio to the
standard unmodified case is
D E
M
L f
D E ≃ f 1− x f x − α = f 1− α . (57)
M
L std

For a change in all masses by a factor f , we obtain the same dependence as


by the individual stellar expression (m/ℓ) ∼ m1−α . We recall that the ratio of the total
mass from lensing to the stellar mass from photometry MMTot∗
was equal to about 5.9 by
van Dokkum [67] and about 2.7 by Mercier et al. [69]. For of factor f = 0.62 (or 0.614 by
Symmetry 2024, 1, 0 21 of 25

[69]) and α = 4.56, we obtain a value of 5.48 (or 5.67) for the ratio in Equation (57). We
obtain, respectively, 4.32 and 2.65 with the redshift by Mercier et al. for exponents α = 4
and 3 in the M-L relation. Such a value α = 3 applies if some star formation is still active
in JWST-ER1, which is not unlikely since this galaxy is observed close to the peak of star
formation in the Universe. Thus, whatever the exact average α-value in JWST-ER1 is, we
may conclude the following: due to the shift in masses, the real scale-invariant M/L ratio is
higher than the one usually considered, which then leads to too low photometric masses, as currently
obtained. In the scale-invariant theory, a standard mass distribution can naturally account
for the lensing mass without calling for dark matter.
In the lens, the stellar masses are smaller, then their luminosity is also smaller. If the
M-L relation were linear, there would be no particular effect for the derived masses, see
(57). But, as m/ℓ is much larger for reduced m, not accounting that is leading to too low
mass estimates.

10
L/LO .

x(log m)
IMF
M-L Super-Salpeter
1000 1

Salpeter

100 Chabrier 0.1

10 0.01

1 .001

30 10 3 1 0.3 M/MO .
Figure 6. The red thick curve represents the mass–luminosity relation on the zero-age sequence
[71], according to the scale indicated on the left vertical axis. The other curves describe various IMF:
Chabrier [68], Salpeter with x = 1.35 [73], and the so-called Super-Salpeter by [67], according to the
scale indicated on the right vertical axis. Note that, in ξ (log m), the log is a decimal. The Figure is
inspired by [67].

6.4. The Masses from the Sloan Lens ACS (SLACS) Survey
To base our analysis on a larger sample, we examine the Sloan Lens ACS (SLACS)
survey by Bolton [74], which is one of the largest and most homogeneous datasets of galaxy-
scale lenses, containing about 100 lenses analyzed by Auger et al. [75]. The distribution of
luminous and dark matter is studied to redshift z ≈ 0.5. It is based on multi-band imaging
with ACS, WFC, and NICMOS on the HST, analyzed with a lens model using an isothermal
ellipsoid mass distribution, and providing high precision mass measurements. The mass
estimates appear unbiased compared to the estimates from SDSS photometry and are in
the range 1010.5 < M < 1011.8 M⊙ , with a typical statistical error of 0.1 dex. The mean
M-L ratio of the sample is ⟨ M/L⟩ = 8.7, corresponding to a mean stellar mass of 1.84 M⊙ ,
which is not much different from the case of JWST-ER1.
Sal p
The survey provides for each lens the fractions f ∗Chab
,Ein and f ∗,Ein of the stellar mass
within the Einstein radius for the Chabrier and Salpeter IMFs [68,73], with respect to the
lensing mass. These fractions are independent of any priors from lensing, so that, in some
Symmetry 2024, 1, 0 22 of 25

cases, fractions higher than 1 were obtained [75]. The mean stellar mass fraction within the
Einstein radius is ⟨ f ∗Chab
,Ein ⟩ = 0.4 ±0.1 for the Chabrier IMF; for the Salpeter’s IMF, the mean
Sal p
is ⟨ f ∗,Ein ⟩ = 0.7 ± 0.2.
In the sample by [75], we select three sub-samples: low redshifts, lenses with
zL ≤ 0.15; intermediate, with z between 0.15 and 0.30; higher redshifts, with zL ≥ 0.30.
This is a compromise between the need to have different mean zL and a sufficient number
of galaxies in each sub-sample. The results are shown in Table 1. We see that the stellar
mass fractions are lower for the higher z-interval: compared to the lowest z-interval, the
mean differences are ∆ f ∗ = 0.092 for the Chabrier IMF and ∆ f ∗ = 0.166 for the Salpeter
IMF. The values of the stellar mass fractions for the intermediate zone lie in between. These
differences are of the same order as the uncertainties. The standard deviations of the
different fractions are also given in Table 1, allowing tests of significance.

Table 1. Analysis of the SLACS Survey: the mean stellar mass fractions within the Einstein radius,
Sal p
⟨ f ∗Chab
,Ein ⟩ and ⟨ f ∗,Ein ⟩ are given for Chabrier and Salpeter IMFs respectively as well as the corresponding
Sal p
standard deviations. The column ⟨ f SIV,Ein ⟩ gives the theoretical value of the stellar mass fraction
with Salpeter’s IMF obtained after accounting for the scale invariant correction; ideally it should be
1.0. The last colum gives the number of galaxies in each sub-sample.
.
Sal p Sal p Sal p
Redshift interval ⟨z⟩ ⟨ f ∗Chab
,Ein ⟩ ⟨ f ∗,Ein ⟩ σ( f ∗Chab
,Ein ) σ ( f ∗,Ein ) ⟨ f SIV,Ein ⟩ N
0 ≤ z ≤ 0.15 0.114 0.449 0.792 0.110 0.197 0.95 17
0.15 ≤ z ≤ 0.30 0.218 0.419 0.745 0.053 0.130 1.04 39
0.30 ≤ z ≤ 0.50 0.383 0.357 0.626 0.079 0.136 1.07 10

Combining the dispersions for the low and high z-intervals, we obtain σ =
0.135 for the difference ∆ f ∗ = 0.449 − 0.357 = 0.092 of the ⟨ f ∗Chab ,Ein ⟩ values. For a normal
distribution, this corresponds to a probability of 50.4% that the difference is significant. For
Sal p
⟨ f ∗,Ein ⟩, the difference is ∆ f ∗ = 0.792 − 0.626 = 0.166 and σ = 0.239, giving a probability of
51.2%. Thus, the differences are possible, but also compatible with a random realization.
The ages t corresponding to the mean redshifts z = 0.114, 0.218, and 0.383 of the samples,
respectively, are t = 0.8591, 0.9113, and 0.9502 in the dimensionless time scale, again for
Ωm = 0.05. According to the modified conservation laws [1], masses are scaling like
t, which implies an increase in the mass with decreasing redshifts. The observations
Sal p
effectively suggest an increase in the fractions f ∗Chab ,Ein and f ∗,Ein . Let us make a comparison
of observations and theory, even if the statistical uncertainties remain large.
As for JWST-ER1, the Salpeter IMF gives masses in better agreement with observa-
tions than Chabrier’s IMF: the mass estimates from photometry at ⟨z⟩ = 0.114 give a stellar
mass equal to 79.2% of the lensing mass. If we account—as for JWST-ER1—for the shift
of the Salpeter IMF resulting from smaller masses, we need to rescale the stellar fraction
in the Einstein radius following Equation (57). For the low redshift bin, we obtain today’s
Sal p
corresponding value of the stellar mass fraction of ⟨ f SIV ⟩ = 0.792 × 0.9502−3.56 = 0.95. For
Sal p
the intermediate and high reshift bins, we obtain the fractions ⟨ f SIV ⟩ = 1.04 and 1.07. These
fractions, close to 1.0, mean that, when a shift in mass due to scale invariance is accounted
for in absence of dark matter, the stellar masses from photometry within the Einstein ring
for a standard IMF are remarkably close to the total masses from lensing. Thus, it releases
or even suppresses the need for dark matter in explaining the discrepancy between stellar
and lensing masses.

7. Conclusions
The SIV theory rests on a simple fundamental principle: enlarging the group of
invariances that subtend the theory of gravitation, as suggested by Dirac [2], by including
scale-invariant properties. Effects occur only in low-density (ϱ < ϱc ) Universe models
Symmetry 2024, 1, 0 23 of 25

and generally depend on the inverse of the age of the Universe. Even in a low-density
Universe, the SIV effects are largely dominated by Newton gravity, while their cumulative
contributions may play some role in the evolution of locally low-density structures. Of note
is that the theory contains no “adjustment parameter”.
There is an impressive convergence of results concerning several fundamental
problems in astrophysics. They include the following: (1) the observed excesses in velocity
for member galaxies in clusters with respect to the visible matter; (2) the relatively flat
or shallow slope of the rotation curves in low redshift galaxies; (3) the observed steeper
rotation curves of high-redshift galaxies; (4) the omnipresent relation between dark and
baryonic matters observed in galaxy clusters and galaxies at different redshifts; (5) the
excess of velocity in very wide binary stars with separations larger than about 3000 kau;
(6) the discrepancy between the (larger) masses of galaxies derived from lensing and the
(lower) stellar masses obtained from their photometry.
The results (with an appropriate size!) all support the occurrence of scale-
invariant effects and throw into question the need for—or even the very existence of—dark
matter. The observations considered here, in addition to those of previous studies (in
particular on the growth of density structures [7]), concern a variety of problems (see
also the introduction); they come from a broad range of sources and authors and they
happen in astronomical objects spanning a vast range of evolutionary stages, with masses
from 1 M⊙ to about 1014 M⊙ , and spatial scales from 0.01 pc to a few Gpc. In view of the
different tests performed, it seems unlikely that the scale-invariant theory explains so many
observations—and ones which are so diverse—just by chance, especially in the absence of
any fitting parameter.

Author Contributions: Writing —original draft A.M.; conceptualization—both authors A.M. and F.C.;
formal analysis—both authors; investigation—both authors; methodology—both authors; validation—
both authors; writing—review and editing—both authors. Both authors have read and agreed to the
published version of the manuscript.
Funding: This research received no external funding.
Data Availability Statement: No new data were created or analyzed in this study.
Acknowledgments: A.M. expresses his gratitude to James Lequeux and Georges Meynet for their
support and encouragement, and to Vesselin Gueorguiev for constructive collaboration..
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Maeder, A. An Alternative to the LambdaCDM Model: The Case of Scale Invariance. Astrophys. J. 2017, 834, 194.
2. Dirac, P.A.M. Long range forces and broken symmetries. Proc. R. Soc. Lond. Ser. A 1973, 333, 403.
3. Canuto, V.; Adams, P.J.; Hsieh, S.-H.; Tsiang, E. Scale-covariant theory of gravitation and astrophysical applications. Phys. Rev. D.
1977, 16, 1643.
4. Maeder, A.; Gueorguiev, V.G.; Action Principle for Scale Invariance and Applications (Part I). Symmetry 2023, 15, 1966.
5. Gueorguiev, V.; Maeder, A. The Scale-Invariant Vacuum Paradigm: Main Results and Current Progress Review (Part II). Symmetry
2024, 16, 657.
6. Maeder, A.; Gueorguiev, V.G. The Scale-Invariant Vacuum (SIV) Theory: A Possible Origin of Dark Matter and Dark Energy.
Universe 2020, 6, 46.
7. Maeder, A.; Gueorguiev, V.G. The growth of the density fluctuations in the scale-invariant vacuum theory. Phys. Dark Universe
2019, 25, 100315.
8. Peterson, M.A. Galileo’s Discovery of Scaling Laws. arXiv 2001, arXiv:00110031.
9. Feynman, R.P.; Leigfhton, R.; Sands, M. Feynman Lectures on Physics, Volume 1: Mainly Mechanics, Radiation and Heat; 1963;
Addison-Wesley, Boston, USA.560p.
10. Weyl, H. Raum, Zeit, Materie. Vorlesungen über allgemeine Relativitätstheorie; Re-edited 1970; Springer Verlag: Berlin/Heidelberg,
Germany, 1923.
11. Eddington, A.S. The Mathematical Theory of Relativity; Chelsea Publ. Co.: New York, NY, USA, 1923; 266p.
12. Einstein, A. Kommentar zu “Hermann Weyl: Gravitation und Elektrizität”; Sitzung Berichte der Königlich Preussischen Akademie
des Wissenschaften: Berling, Germany, 1918; p. 478.
Symmetry 2024, 1, 0 24 of 25

13. Karananas, G.K.; Shaposhnikov, M.; Shkerin, A.; Zell, S. Scale and Weyl invariance in Einstein–Cartan gravity. Phys. Rev. D 2021,
104, 124014.
14. Einasto, J.; Einasto, M.; Gramann, M. Structure nd formation of superclusters—IX. Self-similarity of voids. MNRAS 1989, 238, 155.
15. Bouvier, P.; Maeder, A. Consistency of Weyl’s Geometry as a Framework for Gravitation. Astrophys. Space Sci. 1978, 54, 497.
16. Fujii, Y.; Maeda, K.-I. The Scalar-Tensor Theory of Gravity; Cambridge University Press: Cambridge, UK, 2003.
17. Maeder, A. MOND as a peculiar case of the SIV theory. MNRAS 2023, 520, 1447.
18. Einstein, A. Autobiographical Notes; Open Court Publ. Co.: La Salle: Chicago, IL, USA; 1949.
19. Dirac, P.A.M. Cosmological Models and the Large Numbers Hypothesis. Proc. R. Soc. Lond. Ser. A 1974, 338, 439.
20. Carroll, S.M.; Press, W.H.; Turner, E.L. The cosmological constant. Annu. Rev. Astron. Astrophs. 1992, 30, 499.
21. Maeder, A.; Gueorguiev, V.G. Scale invariance, horizons, and inflation. MNRAS 2021, 504, 4005.
22. Jesus, J.F. Exact solution for flat scale-invariant cosmology. Rev. Mex. Astron. Astrophys. 2018, 55, 17.
23. Aghanim, N.; Akrami, Y.; Ashdown, M.; Baccigalupi, C.; Ballardini, M. Planck 2018 results. VI. Cosmological parameters. Astron.
Astrophys. 2020, 641, A6.
24. Maeder, A.; Bouvier, P. Scale invariance, metrical connection and the motions of astronomical bodies. Astron. Astrophys. 1979, 73, 82.
25. Maeder, A. Dynamical Effects of the Scale Invariance of the Empty Space: The Fall of Dark Matter? Astrophys. J. 2017, 849, 158.
26. Maeder, A. Four Basic Solar and Stellar Tests of Cosmologies with Variable Past G and Macroscopic Masses. A&A 1977, 56, 359.
27. Clifton, T.; Ferreira, P.G.; Padilla, A.; Skordis, C. Modified gravity and cosmology,Phys. Rep. 2012, 513, 1.
28. Ikeda, S.; Saridakis, E. N.; Stavrinos, P.C. Cosmology of Lorentz fiber-bundle induced scalar-tensor theories. Phys. Rev D. 2019,
100, 12403.
29. Zwicky, F. Die Rotverschiebung von extragalaktischen Nebeln. Helevetica Phys. Acta 1933, 6, 10.
30. Karachentsev, I.D. Astrofizika 1966 2, 81. The virial mass-luminosity ratio and the instability of different galactic systems,
Astrophysics 1967, 2, 39. (In English)
31. Sohn, J. Geller , M.J.; Zahid, H.J. et al. The Velocity Dispersion Function of Very Massive Galaxy Clusters: Abell 2029 and Coma.
ApJ Suppl. 2016, 229, 20.
32. Bahcall, N.The Perseus Cluster: Galaxy Distribution, Anisotropy, and the Mass/Luminosity Ratio. ApJ 1974, 187, 439.
33. Blindert, K.; Yee, H.K.C.; Gladders, M.D.; Yee, H K.C.; Gladders, M.D.; Ellingson, E. Dynamical masses of RCS galaxy clusters.
IAU Coll. 2004, 195, 215.
34. Proctor, R.N.; Mendes de Oliveira, C.; Azanha, L. A derivation of masses and total luminosities of galaxy groups and clusters in
the maxBCG catalogue. MNRAS 2015, 449, 2345.
35. Chiu, I.; Mohr, C.L.; McDonald, M.; Bocquet, S.; Ashby, M.L.N.; Bayliss, M. Baryon content of massive galaxy clusters at 0.57 < z <
1.33. MNRAS 2016, 455, 258.
36. Ge, C.; Wang, Q.D.; Tripp, T.M.; Li, Z.; Gu, Q.; Ji, L. Baryon content and dynamic state of galaxy clusters: XMM-Newton
observations of A1095 and A1926. MNRAS 2016, 459, 366.
37. Gonzalez, A.H.; Sivanandam, S.; Zabludoff, A.I. et al. Tracing Cosmic Evolution with Clusters of Galaxies. Available online:
http://www.sexten-cfa.eu/en/conferences/2013/details/34-SestoClusters2013.html accessed on 1 -5 July 2013.
38. Leauthaud, A.; Leauthaud, A.; George, M.R.; Behroozi, P.S.; Bundy, K.; Tinker, J.; Wechsler, R.H.The Integrated Stellar Content of
Dark Matter Halos. ApJ 2012, 746, 95.
39. Lin, Y.-T.; Stanford, S.A.; Eisenhardt, P.R.M.; Vikhlinin, A.; Maughan, B.J.; Kravtsov, A. Baryon Content of Massive Galaxy
Clusters at z = 0 − 0.6. ApJ 2012, 745, L3.
40. Shan, Y.; McDonald, M.; Courteau, S. Revised Mass-to-light Ratios for Nearby Galaxy Groups and Clusters. ApJ 2015, 800, 122.
41. Chan, M. H. Two mysterious universal dark matter-baryon relations in galaxies and galaxy clusters. Phys. Dark Universe 2022, 38,
101142.
42. Trimble, V. Existence and nature of dark matter in the Universe. Annu. Rev. Astron. Astrophs. 1987, 25, 425.
43. Sofue, Y.; Rubin, V. Rotation Curves of Spiral Galaxies. Annu. Rev. Astron. Astrophys. 2001, 39, 137.
44. Huang, Y.; Lin, X.-W.; Yuan, H.-B.; Xiang, M.S.; Zhang, H.W.; Chen, B.Q. The Milky Way’s rotation curve out to 100 kpc and its
constraint on the Galactic mass distribution. MNRAS 2016, 463, 2623.
45. Eilers, A.-C.; Hogg, D.W.; Ris, H.W.; Ness, M. K. The circular velocity curve of the Milky Way from 5 to 25 kpc. Astrophys. J. 2019,
871, 120.
46. Jiao, Y.; Hammer, F.; Wang, H.; Wang, J.; Amram, P.; Chemin, L.; Yang, Y. Detection of the Keplerian decline in the Milky Way
rotation curve. Astron. Astrophys. 2023, 678, A208.
47. Maeder, A. Observational tests in scale invariance I: Galaxy clusters and rotation of galaxies. submitted. arXiv 2024,
arXiv:2403.08759.
48. Genzel, R.; Forster Schreiber, N.M.; Ubler, H. Strongly baryon-dominated disk galaxies at the peak of galaxy formation ten billion
years ago. Nature 2017, 543, 397.
49. Lang, P.; Forster Schreiber, N.M.; Genzel, R. Falling Outer Rotation Curves of Star-forming Galaxies at 0 <z < 2.6 Probed with
KMOS3D and SINS/zC-SINF. ApJ 2017, 840, 92L.
50. Genzel, R.; Price, S. H.; Ubler, H.; Schreiber, N.F.; Shimizu, T.T.; Tacconi, L.J. Rotation Curves in z 1–2 Star-forming Disks: Evidence
for Cored Dark Matter Distributions. ApJ 2020, 902, 98.
Symmetry 2024, 1, 0 25 of 25

51. Nestor Shachar, A.; Price, S.H.; Schreiber, N.F.; Genzel, R.; Shimizu, T.T.; Tacconi, L.J. RC100: Rotation Curves of 100 Massive
Star-forming Galaxies at z = 0.6 − 2.5 Reveal Little Dark Matter on Galactic Scales. ApJ 2023, 944, 78.
52. Lelli, F.; McGaugh, S.S.; Schombert, J.M.; Pawlowski, M.S. One Law to Rule Them All: The Radial Acceleration Relation of
Galaxies. ApJ 2017, 836, 152.
53. Maeder, A.; Gueorguiev, V.G. Scale-invariant dynamics of galaxies, MOND, dark matter, and the dwarf spheroidal. MNRAS 2020,
492, 2698.
54. Hernandez, X.; Cookson, S.; Cortes, R.A.A. Internal kinematics of Gaia eDR3 wide binaries. MNRAS 2022, 502, 2304.
55. Brown, A.G.; Vallenari, A.; Prusti, T.; De Bruijne, J.H.; Babusiaux, C.; Biermann, M. Gaia Early Data Release 3. The Gaia Catalogue
of Nearby Stars. Astron. Astrophys. 2021, 649, A1.
56. Milgrom, M. A modification of the newtonian dynamics: Implications for galaxy systems. Astrophys. J. 1983, 270, 365.
57. Milgrom, M. The Mond Limit from Spacetime Scale Invariance. Astrophys. J. 2009, 698, 1630.
58. Hernandez, X. Internal kinematics of Gaia DR3 wide binaries: Anomalous behaviour in the low acceleration regime. MNRAS
2023, 525, 1401.
59. Pittordis, C.; Sutherland, W. Wide Binaries from Gaia EDR3: Preference for GR over Mond ? Open J. Astrophys. 2023, 6, 4.
60. Chae, H.-K. Breakdown of the Newton—Einstein Standard Gravity at Low Acceleration in Internal Dynamics of Wide Binary
Stars. ApJ 2023, 952, 128.
61. Banik, I.; Pittordis, C.; Sutherland, W.; Famaey, B.; Ibata, R.; Mieske, S.; Zhao, H. Strong constraints on the gravitational law from
Gaia DR3 wide binaries. MNRAS 2024, 527, 4573.
62. Chae, H.-K. Robust Evidence for the Breakdown of Standard Gravity at Low Acceleration from Statistically Pure Binaries Free of
Hidden Companions. arXiv 2024, arXiv:2309.10404.
63. El-Badry, K.; Rix, H.W.; Heinz, T.M. A million binaries from Gaia eDR3: Sample selection and validation of Gaia parallax
uncertainties. MNRAS 2269, 506.
64. Jiang, Y.-F.; Tremaine, S. The evolution of wide binary starsThe evolution of wide binary stars. MNRAS 2010, 401, 977.
65. Loeb, A. Gravitational Redshift for Wide Binaries in Gaia eDR3. Res. Note AAS 2022, 6, 55.
66. Kottler, F. Uber die physikalischen Grundlagen der Einsteinschen Gravitationstheorie. Ann. Phys. 1918, 361, 401.
67. Van Dokkum, P.; Brammer, G.; Wang, B.; Leja, J.; Conroy, C. A massive compact quiescent galaxy at z = 2 with a complete Einstein
ring in JWST imaging. Nat. Astron. 2023, 8, 119.
68. Chabrier, G. Galactic Stellar and Substellar Initial Mass Function. Publ. Astron. Soc. Pac. 2003, 115, 763.
69. Mercier, W.; Shuntov, R.; Gavazzi, J.W.; Nightingale, J.W.; Arango, R.; Ilbert, O. The COSMOS-Web ring: In-depth characterization
of an Einstein ring lensing system at z ∼ 2. A&A 2024, 687, A61.
70. Johnson, B.D.; Leja, J.; Conroy, C.; Speagle, J.S. Stellar Population Inference with Prospector. ApJ Suppl. Ser. 2021, 254, 22.
71. Maeder, A. Physics, Formation and Evolution of Rotating Stars; Springer Verlag: Berlin/Heidelberg, Germany, 2009; 829p.
72. Longair, M.S. Galaxy Formation; Springer: Berlin/Heidelberg, Germany, 1998; 536p.
73. Salpeter, E. The Luminosity Function and Stellar Evolution. ApJ 1955, 121, 161.
74. Bolton, A.; Burles, S.; Koopmans, L.V.; Treu, T.; Gavazzi, R.; Moustakas, L.A. The Sloan Lens ACS Survey. V. The Full ACS
Strong-Lens Sample. ApJ 2008, 682, 964.
75. Auger, M.W.; Treu, T.; Bolton, A.S.; Gavazzi, R.; Koopmans, L.V.E.; Marshall, P.J. The SLOAN lens ACS Survey. XI. Colors, lensing
and stellar masses of early-type galaxies. ApJ 2009, 705, 1099.

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like