The alert reader must have anticipated the main result of the previous section, namely, that consists of freely moving particles. By linearity, particles can move through one another uninterrupted and if so, they are noninteracting particles which should better have straight paths. Enabling their mutual interaction therefore requires some form of nonlinearity, either in the coarsener, , or in the scaling part. Further recalling our commitment to general covariance as a precondition for any fundamental physical theory, nonlinearity is inevitably and, in a sense, uniquely forced upon us. A nonlinear model also supports a plurality of particles, having different sizes which are different from the common in a linear theory. This frees , ultimately estimated at km, to play a role at astrophysical scales.
Operating with
on (
16), the second term of this operator almost annihilates the coarsener by the above remarks (charge conservation), leaving only a curvature term
coming from commutator (
20). The scaling piece combines the covariant generalization of the flat spacetime conversion
with a novel nonlinear term (see
Sec. 3.2 below). The l.h.s. reads
. We would like to swap the order of
and
, which would give
by (
19). However, in curved spacetime
could implicitly depend on the scale
s through
, introducing a commutator correction. The combined result reads
more suited for analysis. For example, in the case of flat spacetime and naive scaling,
, the particle solution of the
J-field associated with (
18) is the familiar Gaussian
from the previous sections. It is emphasized that
and its associated scale flow are merely analytic tools, not to be put on equal footing with
and its flow (As the nonlinear terms arising from scaling and commutator do not involve
but rather
, for a given
and
, (
22) describes the linear but inhomogeneous—in both spacetime and scale—flow of
). Moreover, infinitely many other markers of an
A-particle’s center exist, involving, e.g., curvature terms, which may turn out to be more suitable in certain domains. However, for the purpose of this introductory paper, focusing on weak gravitational fields/flat-spacetime,
turns out to be a most useful choice.
Relation (
19) is formally equivalent to Maxwell’s equations with
sourcing
’s wave equation. However,
is not an independent object as in classical electrodynamics but a marker of the locus of privileged points at which the Maxwell coarsener does not annihilate
(distinct
’s differing by some
therefore have identical
’s). For
and
to mimic those of classical electrodynamics,
must also be localized along curved worldlines traced by solutions of the Lorentz force equation in
(which as already shown in the scalar case, necessitates a nonlinear scale flow). And just like in the scalar-particle case, where higher order cumulants (
) are ‘awakened’ by its center’s nonuniform motion, deforming its stationary shape, so does the
“adjunct" (in the jargon of action-at-a-distance electrodynamics) to each such
gets deformed. Due to the extended nature of an
A-particle, and unlike in
models
3, these deformations at
are
not encoded in the local motion of its center at time
t, but rather on its motion at retarded and advanced times,
(assuming flat spacetime for simplicity). However, associating such temporal incongruity with ‘radiation’ can be misleading, as it normally implies the freedom to add any homogeneous solution of Maxwell’s equations to
which is clearly nonsensical from our perspective. Consequently, the retarded solution cannot be imposed on
and in general,
contains a mixture of both advanced and retarded parts, which varies across spacetime. The so-called radiation arrow of time manifested in every macroscopic phenomenon must therefore receive an alternative explanation (see
Section 3.4.2).
Now, why should
be confined to the neighborhood of a worldline? As already seen in the linear, time-dependent case, a scale-flow such as (
16) suffers from instability in both
s-directions due to the action of the coarsener; Local variations of
along space-like/time-like directions are rapidly amplified in the
/
direction resp. If we therefore examine the scale flow of
inside a ‘lab’ of dimension much smaller than
, centered at the origin without loss of generality, then the coarsener would completely dominate the flow of
whose scale of variation is much smaller than
(unlike (
18)!) leading to its rapid divergence, unless it is
almost annihilated by
, i.e.,
varies exclusively in light-like directions. An exception to this rule could take place around privileged points, where the scaling-field grows to such large magnitude as to allow for the scaling piece to counter a large coarsener contribution. This is where
is focused and, as shown in
Section 3.1.2 below, it is only around world-lines marking the center of an extended particle, where
peaks, that
can indeed grow to such large magnitude. “Almost" is emphasized above because it is precisely the fact that, at distances from
that are much smaller than
, the action of the coarsener is only reduced to the order of that of the scaling piece, which gives rise to deviations from classical physics. This will be a recurrent theme in the rest of the paper.
3.2. The Motion of Matter Lumps in a Weak Gravitational Field
Equation (
48) prescribes the scale flow of the metric (in the Newtonian approximation), given the a set,
, of worldlines associated with matter lumps. To determine this set, an equation for each
, given
, is obtained in this section. This is done by analyzing the scale flow of the first moment of
associated with a general matter lump, using (
22). An obstacle to doing so comes from the fact that,
now incorporates both gravitational and non gravitational interactions in a convoluted way, as the existence of gravitating matter depends on it being composed of charged matter. In order to isolate the effect of gravity on
, we first analyze the motion of a body in the absence of gravity, i.e.,
,
with
the Minkowskian coordinates and
the structural component from the previous sections. To this end, a better understanding of the scale flow (
22) of
is needed. Using (
17) and (
19) plus some algebra the scaling piece in (
22) reads
with
(where
is of course only defined up to a divergence free piece). The first two terms in (
49) are the familiar
conversion, to which a ‘matter vector’,
, is added. In the absence of gravity the commutator on the l.h.s. of (
22) vanishes. Combined, we get
Since
, taking the divergence of (
51) implies
where the second equality follows from the Maxwell-Bianchi identity. This constraint, relating
to
inside matter, is the counterpart of (
32), derived this time from the flow of
rather than that of
, and from local charge conservation rather than local energy-momentum conservation; Note that our particle solution from
Section 3.1.2 automatically satisfies (
52) by virtue of
. In the case of a general
, operating with
on (
22), the l.h.s vanishes identically as it is just the covariant divergence of
(
19) with
there. Using (
20) the r.h.s gives a covariant counterparts of (
52) involving curvature terms, but still expressing a section constraint.
Defining
and
the (conserved in time-) electric and ‘matter’ charges resp., and integrating (
51) over three-space implies
i.e., electric charge is conserved in scale if and only if the matter charge vanishes. That the latter is identically true follows from the divergence form (
50) of
and
. Whether this result generalizes to curved spacetime is unclear to the author at this point
4. Clearly, since charge is conserve in time,
whenever spacetime is approximately locally flat around
some point on the world-line of a charge, but it is possible that a center more complicated than
(
19) is required for identical charge conservation in scale. Alternatively, charge may not be identically conserve in scale when strong gravitational fields are involved, or it could emerge as a necessary condition for inclusion in
, as in the linear model.
Next, multiplying (
51) by
and integrating over a ball,
B, containing a body of charge
q, results in
where
is an object’s ‘center-of-charge’ (c.o.c.). Above and in the rest of this section, the charge of a body, assumed nonzero for simplicity, is only used as a convenient tracer of matter. Now, as
can be both positive and negative, the c.o.c. is not necessarily confined to the support of
, as with positive distributions. Nonetheless, since
,
does follow the particle up to some constant displacement, reflecting to a large extent the arbitrariness in defining the exact position of an extended body, and assumed much smaller than any competing length.
At the (sub-)atomic level,
would be rapidly fluctuating in time, endowing
with a ‘jitter motion’ component. To remove it, (
53) is convolved with a normalized symmetric kernel of macroscopic extent
T. Defining
the form of (
53) is retained for
(with
), except for a correction
coming from the time-scaling term. If
is slowly varying over
T, this correction is at most on the order of
, negligibly renormalizing the
term for
. For economical notations, then, the ‘bar’ is dropped henceforth from all quantities.
The integral in (
53) is the first moment of a distribution,
, whose zeroth moment vanishes, which is therefore invariant under
—as is expected of a ‘force term’, competing with the
, acceleration term. Buried in it are presumably all forms of non gravitational interactions preventing
for some constant
and
, from being a solution of (
53) in the absence of gravity (In particular, elements of the Lorentz force can explicitly be excavated from it, but this route shall not be explore here because the Lorentz force is implicit in the basic tenets of classical electrodynamics and, moreover, as there is only a single, global
A-field, its decomposition into
, an adjunct ‘self-field’, and an external field, is far from being obvious). Accordingly, this term is ignored for an isolated body, for which all forces are internal and necessarily integrate to zero (this can be verified explicitly, e.g., for a spherically symmetric
). Thus (
53) becomes (
14), and as proved in that case, solutions for
must all be straight, non tachyonic worldlines
.
The effect of gravity on those straight worldlines is derived by including a weak field in the flow of the first-moment projection of (
22). Recalling from
Section 3.1.2 that weak gravity has no effect on the scaling field, and since gravity is assumed to play a negligible role in the structure of matter, the way this field enters the flat spacetime analysis is by ‘dotting the commas’ in partial derivatives. Before analyzing the first moment projection we first have to show that the identically conserved-in-time charge by virtue of the last identity in (
19),
is also conserved in scale. As previously remarked, this might automatically follow from the previously mentioned covariant counterpart of (
52), but since that more complicated constraint had not been explored, and as a preparation for the first-moment analysis, we explicitly show next that (
54) hold up to
corrections. To this end (
22) is evaluated using the Newtonian metric (
40), then multiplied by
and integrated over three-space. A straightforward calculation to first order in
, incorporating
, gives
Ignoring
corrections to the isotropic coarsener, the net effect of of the potential in the Newtonian approximation is to render the coarsener anisotropic through its gradient, with an added relativistic correction in the form of the last term on the r.h.s. of (
55). The reader can then verify that the contribution of (
55) vanishes up to
corrections (see also below, the corresponding contribution to the first moment projection). The same is true for the contribution of the scaling piece, since
for weak fields and
. As for the contribution of the non-vanishing commutator on the l.h.s. of (
22), using
(viz., ordinary derivatives can replace covariant ones) in the definition (
19) of
, we have
Following integration by parts of the second term, the contribution of the commutator, combined with that of , reads . Finally, since , the contribution of the curvature term vanishes, as is that of the covariant term up to corrections.
Moving to the first-moment projection of (
22), we multiply it by
, plug expression (
55) for the d’Alembertian and integrate over
B, assuming
is approximately constant over the extent of the body. At non-relativistic velocities and neglecting
corrections, the double time-derivative piece contributes
, where
. In the second term on the r.h.s. of (
55) the
cancels (to first order in
) the
factor multiplying it, which is then integrated by parts. Neglecting
corrections the result is
. Using the continuity equation for
, integration by parts of the last term in (
55) yields a negligible
, vanishing at non-relativistic velocities. As with the zeroth-moment project, the modification to the scaling piece (
49) only introduces an
correction to the
term in (
53) which is neglected in the Newtonian approximation. Finally, the contribution
, combined with the commutator (
56), gives on the l.h.s.:
. This last integral is evaluated by multiplying
by
and integrating by parts the l.h.s., which only modifies the
in the first piece.
Combining all pieces, the first moment projection of (
22) reads
This equation is just (
14) with an extra ‘force-term’ on its r.h.s. which could salvage a non uniformly moving solution,
, from the catastrophic fate at
suffered by its linear counterpart.
At sufficiently large scales,
, when all relevant masses contributing to
occupy a small ball of radius
centered at the origin of scaling without loss of generality (
can similarly be assumed) the scaling part on the r.h.s of (
57) becomes negligible compared with the force term, rather benefiting from such crowdedness. It follows that each
would grow—extremely rapidly as we show next—with increasing
even when the weak-field approximation is still valid, implying that the underlying
is not in
. The only way to keep the scale evolution of
under control is for the acceleration term to similarly grow,
almost canceling the force term but not quite, which is critically important; it is the fact that the sum of these two terms, both originating from the coarsener, remains on the order of the scaling term, which is responsible for a nontrivial, non pure scaling
. This means that each worldline converges at large scales to that satisfying Newton’s equation
At small scales the opposite is true. The scaling part dominates and any
scaling path, i.e.,
is well behaved. Combined: at large scales
is determined, then simply scaled at small scales, gradually converging locally to a freely moving particle. Finally, (
58) must be true also for a loosely bound system, e.g. a wide binary moving in a strong external field, implying that Newton’s law applies also to their relative vector, not merely to their c.o.m.
Deriving a manifestly covariant generalization of (
57) is certainly a worthwhile exercise. However, in a weak field the result could only be
with
some scalar parameterization of the worldline traced by
, and
the gravitational part of the scaling field, well-approximated by
. Above,
are the Christoffel symbols associated with
, i.e., the analytic continuation of the metric, seen as a function of Newton’s constant, to
. Recalling from
Section 3.1.3 that the fixed-point
is a solution of the standard EFE analytically continued to
,
in that case is therefore just the Christoffel symbol associated with standard solutions of EFE’s. The previous, Newtonian approximation is a private case of this, where
contains a factor of
G. Note however that the path of a particle in our model is a covariantly defined object irrespective of the analytic properties of
. Resorting to analyticity simply provides a constructive tool for finding such paths whenever
is analytic in
G. In such cases, the covariant counterpart of (
58) becomes the standard geodesic equation of GR which gives great confidence that this is also the case for non-analytic
.
The reasons for trusting (
59) are the following. It is manifestly scale- and general-covarinat, as is our model; it is
-shift invariant, i.e.,
, parameterizing the same, scale dependent world-line, also solves (
59) for any
; At nonrelativistic velocities in a Minkowskian background,
solves (
59) which, when substituted into the
i-components of (
59), recovers (
57); The scaling regime ansatz,
, solves
, i.e., each point on the world-line traced by
, indexed by a fixed
, flows along integral curves of the scaling field—as must be the case when the coarsener is negligible; It only involves local properties of
and
, i.e., their first two derivatives, which must also be a property of a covariant derivation, as is elucidated by the non-relativistic case. Thus (
59) is the only candidate up to covariant, higher derivatives terms involving
and
, or nonlinear terms in their first or second derivatives, all becoming negligible in weak fields/ at small accelerations.
In conclusion of this section we wish to relate (
59) to the fact that
is not covariantly conserved by virtue of (
28). It was well known already to Einstein that the geodesic equation follows from local energy-momentum conservation under reasonable assumptions. Similarly, the Lorentz force equation follows quite generally from the basic tenets of classical electrodynamics (
31),(
30),(
19). Referring to
Figure 1, both results are derived by integrating
over a world-cylinder,
C, with the
term obtained by converting part of the volume integral into a surface integral over the
’s via (a relativistic generalization of-) Stoke’s theorem, leaving the remaining part for the ‘force term’, which gives
in the case of gravity. The integral over
T represents small radiative corrections to the geodesic/Lorentz-force equation which can be ignored in what follows. A crucial point in that derivation is that the result is insensitive to the form of
C so long as the
’s contain the support of the particle’s energy-momentum distribution. This insensitivity would not carry to the r.h.s. of (
28) should it be transferred sides, whether or not
is to include the self-field of the particle. Thus attempting to generalize the geodesic equation based on (
28), in the hope that it would reproduce fixed-
sections of (
59) solutions, is bound to fail. Nonetheless, since the conservation-violating r.h.s. of (
28) is a tiny
, it can consistently be ignored for any reasonable choice of
C whenever the individual terms in the geodesic equation are much larger, i.e., at large scales; it is only at small scales that each term becomes comparable to the r.h.s. even for reasonable choices of
C. And since in the limit
paths become simple geodesics at any scale according to (
59), the two length parameters,
,
are not independent, constrained by consistency of solutions such as the above, and likely additional ones involving the structure of matter, as hinted to by the fixed-point example in
Section 3.1.2.
3.2.1. Application: The Rotation Curve of Disc Galaxies
As a simple application of (
57), let us calculate the
rotation curve,
, of a scale-invariant mass,
M, located at the origin, as it appears to an astronomer of native scale
. Above,
r is the distance to the origin of a test mass orbiting
M in circles at velocity
v. Since
in (
57) is time-independent, the time-dependence of
can only be through the combination
for some function
. Looking for a circular motion solution in the
plane,
and equating coefficients of
and
for each component, the system (
57) reduces to two, first order ODE’s for
and
. The equation for
readily integrates to
for some integration constant
, and for
r it reads
Solutions of (
60) with
as initial condition, all diverge in magnitude for
except for a single value of
for which
in that limit; for any other
,
rapidly diverges to
respectively; the map
is invertible. We note in advance that, for a mistuned
this divergence starts well before the weak field approximation breaks down due to the
term, and neglected relativistic and self-force terms become important, and being so rapid,
, those would not tame a rogue solution. It follows that there is no need to complicate our hitherto simple analysis in order to conclude that
is a necessary conditions for
r to correspond to an
.
Solutions of (
60) which are well-behaved for
admit a relatively simple analytic form. Reinstating
c and defining
the result is
having the following power law asymptotic forms
with the corresponding asymptotic circular velocity,
With these asymptotic forms the reader can verify that, in the large
regime, (
58), which in this case takes the form:
is indeed satisfied for any
.
It is conjectured that the above result generalizes as follows to the case, of multiple, gravitationally interacting, scale-independent masses
(p a particle index): The smooth,
-dependent family of solutions
to (
57)(
48) for a bound system, approaches
at large scale, where
are Newtonian paths of interacting point-masses
. The scaling form (
65) is an exact symmetry of Newtonian gravity, and since well behaved solutions for the flow (
57)(
48) must approach at large
a
smooth-in-scale family of Newtonian solutions, a solution other than (
65) seems impossible in generic cases.
Finally, for a scale-dependent mass in (
60), an
is obtained by the large-scale regularity condition which is not of the form
. This results in a rotation curve
which is not flat at large
, and an
which, depending on the form of
, may not even converge to zero at large
.
Moving to a more realistic representation of disc galaxies. For a general gravitational potential,
, sourced via (
48) by a planar, axially symmetric mass density
—
being the radial distance from the galactic axis in the galactic plane—and for a test mass circularly orbiting the symmetry axis in the galactic plane, the counterpart of (
60) reads
The time-independent mass density,
, is approximated by a (sufficiently dense) collection of concentric line-rings, each composed of a (sufficiently dense) collection of particles evenly spaced along the circumference. Next, recall that in the above warm-up exercise, the solution of (
66) for each such particle is well behaved only for one, carefully tuned value of
. This sensitivity results form instability of the o.d.e. (
66) in the
direction, inherited from that of
, and is not a peculiarity of the Coulomb potential. To find the rotation curve one needs to simultaneously propagate with (
66) each ring—or rather a single representative particle from each ring—
, using an initial guess for
(where
is now a ring index, labeling the ring whose radius at
equals
, viz.,
). Unlike in the previous case, the (mean-field) Newtonian potential of the disc at scale
,
, solution of (
42), must be re-computed at each
. The rotation curve is obtained as that (unique) guess,
, for which no ring diverges in the limit
. In so finding the rotation curve the scale dependence of individual particles comprising the disc needs to be specified. If those are fixed-point particles then their mass is scale-independent by definition. Moreover, as mentioned above, a scale-dependent mass leads to manifest contradictions with observations. In light of this, a scale independent mass is assumed modulo some caveats discussed in
Section 3.4. Note the implication of the scale-invariant-mass approximation, applicable to any gravitating system: Although
is the desired spacetime path, by construction and the
s-translation invarinace of (
57),
for any
s would also be a permissible path at
. In other words, (
57) with the regularity condition at
, generates a continuous family of spacetime paths which could be observed at any fixed native scale,
in particular.
The algorithm described above for finding the rotation curve, although conceptually straightforward, could be numerically challenging and will be attempted elsewhere (see below a more promising algorithm). However, much can be inferred from it without actually running the code. Mass tracers lying at the outskirts of a disc galaxy experience almost the same,
potential, where
M is the galactic mass, independently of
. This is clearly so at
, as higher order multipoles of the disc are negligible far away from the galactic center, but also at larger
, as all masses comprising the disc converge towards the center, albeit at different paces. The analytic solution (
62) can therefore be used to a good approximation for such traces, implying the following power-law relation between the asymptotic velocity,
, of a galaxy’s rotation curve and its mass,
M,
Such an empirical power law, relating
M and
, is known as the
Baryonic Tully-Fisher Relation (BTFR), and is the subject of much controversy. There is no concensus regarding the conssistency of observations with a zero intrinsic scatter, nor is there an agreemnet about the value of the slope—3 in our case—when plotting
vs.
. Some groups [
5] see a slope
while other [
6] insisting it is closer to 4 (both ‘high quality data’ representatives, using primary distance indicaors). While some of the discrepancy in slope estimates can be attributed to selection bias and different methods of estimating the galactic mass, the most important factor is the inclusion of relatively low-mass galaxies in the latter. When restricting the mass to lie above
, almost all studies support a slope close to 3. The recent study [
7] which includes some new, super heavy galaxies, found a slope
and a
-axis intercept of
for the massive part of the graph. Since the optimization method used in finding those two parameters is somewhat arbitrary, imposing a slope of 3 and fitting for the best intercept is not a crime against statistics. By inspection this gives an intercept of
, consistent with [
5], which by (
67) corresponds to
km.
With an estimate of
at hand, yet another prediction of our model can be put to test, pertaining to the radius at which the rotation curve transitions to its flat part. The form (
62) of
implies that the transition from the scaling to the coarsening regime occurs at
. At that scale the radius assumes a value
, which is also the radius at which the force term equals the scaling term. Using standard units where velocities are given in km/s and distances in kpc, gives
. Now, in galaxies with a well-localized center—a combination of a massive bulge and (exponential) disc—most of the mass is found within a radius
, lying to the right of the Newtonian curve’s maximum. Approximating the potential at
by
, the transition of the rotation curve from scaling to coarsening, with its signature rise from a flat part seen in
Figure 2, is expected to show at
, followed by a convergence to the galaxy-specific Newtonian curve. This is corroborated in all cases—e.g. galaxies NGC2841, NGC3198, NGC2903, NGC6503, UGC02953, UGC05721, UGC08490... in Figure 12 of [
8]
The above sanity checks indicate that the rotation curve predicted by our model cannot fall too far from that observed, at least for massive galaxies; it is guarantied to coincide with the Newtonian curve near the galactic center, depart from it approximately where observed, eventually flattening at the right value. However, these checks do not apply to diffuse, typically gas dominated galaxies, several orders of magnitude lighter. More urgently, a slope
is difficult to reconcile with [
6] which finds a slope
when such diffuse galaxies are included in the sample. Below we therefore point to two features of the proposed model possibly explaining said discrepancy. First, our model predicts that
attributed in [
6] to such galaxies would turn out to be a gross overestimation should their rotation curves be
significantly extended beyond the handful of data points of the flat portion. To see why, consider an alternative,
reverse scheme for finding a rotation curve, also applicable to any bound system:
1. Start with a guess for the mass distribution of a galaxy at some large enough scale,
, such that its rotation curve is fully Newtonian (If our conjecture regarding (
65) is true then the flow to even larger
is guaranteed not to diverge for any such initial guess).
2. Let this Newtonian curve flow via (
66)(
48) to
—no divergence problem in this, stable direction of the flow—comparing the resultant mass distribution at
with the observed distribution
3. Repeat step 1 with an improved guess based on the results of 2, until an agreement is reached.
By construction the solution curve is Newtonian at
, having a
tail past the maximum, whose
rightmost part ultimately evolves into the flat segment at
. We can draw two main distinctions between the flows to
of massive and diffuse galaxies’ rotation curves. First, since the hypothetical Newtonian curve at
—that which is based on baryonic matter only—is rising/leveling at the point of the outmost velocity tracer in the diffuse galaxies of [
6], we can be certain that this tracer was at the rising part/maximum of the
curve, rather than on its
tail as in massive galaxies. This means that, in massive galaxies, the counterpart of the short, flat segment of a diffuse galaxy’s r.c., is rather the short flat segment near its maximum, seen in most such galaxies near the maximum of the hypothetical Newtonian curve. Second, had tracers further away from the center been measured in diffuse galaxies, the true flat part would have been significantly lower relative to this maximum than in massive galaxies. With some work this can be shown via the inhomogeneous flow of
derived from (
66)
where
r is a solution of (
66) (re- expressed as a function of
). The gist of the argument is that, solutions of (
68) deep in the coarsening regime (i.e.,
, with
the enclosed mass at
r) upon flowing to smaller
, decay approximately as
, whereas in the scaling regime they remain constant (see also (
63)). In massive galaxies the entire flow from
to
of a tracer originally at the maximum of the r.c.,
, is in the coarsening regime, while in diffuse galaxies it transitions to a scaling regime.
The second possible explanation for the slopes discrepancy, which could further contribute to an intrinsic scatter around a straight BTFR, involves a hitherto ignored transparent component of the energy-momentum tensor. As emphasized throughout the paper, the
A-field away from a non-uniformly moving particle (almost solving Maxwell’s equations in vacuum) necessarily involves both advanced and retarded radiation. Thus even matter at absolute zero constantly ‘radiates’, with advanced fields compensating for (retarded) radiation loss, thereby facilitating zero-point motion of matter. The
A-field at spacetime point
away from neutral matter is therefore rapidly fluctuating, contributed by all matter at the intersection of its worldline with the light-cone of
. We shall refer to it as the Zero Point Field (ZPF), a name borrowed from Stochastic Electrodynamics although it does not represent the very same object. Being a radiation field, the ZPF envelopes an isolated body with an electromagnetic energy ‘halo’, decaying as the inverse distance squared—which by itself is not integrable!—merging with other halos at large distance. Such ‘isothermal halos’ served as a basis for a ‘transparent matter’ model in a previous work by the author [
2] but in the current context its intensity likely needs to be much smaller to fit observations. Space therefore hosts a non-uniform ZPF peaking where matter is concentrated, in a way which is sensitive to both the type of matter and its density. This sensitivity may result both in an intrinsic scatter of the BTFR, and in a systematic departure from ZPF-free slope=3 at lower mass. Indeed, in heavy galaxies, typically having a dominant massive center, the contribution of the halo to the enclosed mass at
is tiny. Beyond
orbiting masses transition to their scaling regime, minimally influenced by additional increase in the enclosed mass at
r. The situation is radically different in light, diffuse galaxies, where the ratio of
is much higher throughout the galaxy, and much more of the non-integrable tail of the halo contributes to the enclosed mass at the point where velocity tracers transition to their scaling regime. This under estimation of the effective galactic mass, increasing with decreasing baryonic mass, would create an illusion of a BTFR slope greater than 3.
3.2.2. Other Probes of ‘Dark Matter’
Disc galaxies are a fortunate case in which the worldline of a body transitions from scaling to coarsening at a common scale along its entire worldline (albeit different scales for different bodies). They are also the only systems in which the velocity vector can be inferred solely from its projection on the line-of-sight. In pressure supported systems, e.g., globular clusters, elliptical galaxies or galaxy clusters, neither is true. Some segments of a worldline could be deep in their scaling regime while others in the coarsening, rendering the analysis of their collective scale flow more difficult. The above reverse scheme leverages the fact that, all the worldlines of a bound system are deep in their coarsening regime at sufficiently large scale, where their fixed-
dynamics is well approximated using Newtonian gravity. Starting with such a Newtonian system at sufficiently large
, the integration of (
57) to small
is in its stable direction, hence not at risk of exploding for any initial choice of Newtonian paths. If the Newtonian system at
is chosen to be virialized, a ‘catalog’ of solutions of pressure supported systems extending to arbitrarily small
can be generated, and compared with line-of-sight velocity projections of actual systems. As remarked above, the transition from coarsening to scaling generally doesn’t take place at a common scale along the worldline of any single member of the system. However, if we assume that there exists a rough transition scale,
, for the system as a whole in the statistical sense, which is most reasonable in the case of galaxy clusters, then immediate progress can be made. Since in the scaling regime velocities are unaltered, the observed distribution of the line-of-sight velocity projections should remain approximately constant for
, that of a virialized system, viz., Gaussian of dispersion
. On the other hand, at
a virialized system of total mass
M satisfies
where
is the velocity dispersion, and
r is the radius of the system, which is just (
64) with
. On dimensional grounds it then follows that
would be the counterpart of
from (
67), implying
which is in rough agreement with observations. The proportionality constant can’t be exactly pinned using such huristic arguments, but its observed value is on the same order of magnitude as that implied by (
67).
Applying our model to gravitational lensing in the study of dark matter requires better understanding of the nature of radiation. This is murky territory even in conventional physics and in next section initial insight is discussed. To be sure, Maxwell’s equations in vacuum are satisfied away from
, although only ‘almost so’, as discussed in
Section 3. However, treating them as an initial value problem, following a wave-front from emitter to absorber is meaningless for two reasons. First, tiny,
local deviations from Maxwell’s equations could become significant when accumulated over distances on the order of
. Second, in the proposed model extended particles ‘bump into one another’ and their centers jolt as a result—some are said to emit radiation and other absorb it, and an initial-value-problem formulation is, in general, ill-suited for describing such process. Nonetheless, incoming light—call it a photon or a light-ray—does posses an empirical direction when detected. In flat spacetime this could only be the spatial component of the null vector connecting emission and absorption events, as it is the only non arbitrary direction. A simple generalization to curved spacetime, involving multiple, freely falling observers, selects a path,
, everywhere satisfying the
light-cone condition. Every null geodesic satisfies the light-cone condition, but not the converse. In ordinary GR, the only non arbitrary path connecting emission and absorption events which respects the light-cone condition and locally depends on the metric and its first two derivatives is indeed a null geodesic. In our model, a solution of (
59) which is well behaved on all scales, further satisfying the light-cone condition for
, is an appealing candidate: It must (almost) be a null geodesic in strong fields/at large scales, and when transitioning to weak fields/small scales, the light-cone condition inherited from large scales is preserved by the scaling operation; Indeed, denoting
, applying
to both sides of
multiplying the result by
, one gets
In arriving at (
70) use of (
24) has been made in converting
to spacetime derivatives of
and
, and a term
has been neglected (which is justified in the context of cosmology, discussed in
Sec. 3.4.2). Likewise, all spacetime derivatives of
have been neglected (weak field domain). It then follows that
if
.
Furthermore, in GR the deflection angle of a light ray due to gravitational lensing, by a compact gravitating system of mass
M, is
, where
R is the impact parameter. When
is in its scaling regime, our model’s
remains constant. If the system is likewise in its scaling regime, (
69) implies that its virial mass,
, scales according to
, as does the impact parameter of
,
. The conventional mass estimate based on the virial theorem, of this
-dependent family of gravitating systems, would then agree with that which is based on (conventional) gravitational lensing,
, up to a constant, common to all members—recall that this entire family appears in the ‘catalog’ of
systems. Extending this family to large
, the two estimates will coincide by construction. Thus if the system and
transition approximately at a common
, this proportionality constant can only be close to 1. This is apparently the case in most observations pertaining to galaxy clusters. Nonetheless, the two methods according to our model need not, in general, produce identical results. The degree to which they disagree depends on the exact scale-flow of
, which is one of those calculations avoided thus far, involving a path non-uniformly (in scale) transitioning from coarsening to scaling.
A final caveat regarding the application of (
59) is that, we have no reason to expect that the light-cone condition is satisfied
exactly (other than in the
limit). Nonetheless, in and of itself this caveat does not invalidate such solutions, even as candidates for exact solutions, so long as no conflict with observations arises.
3.4. Cosmology
Cosmological models are stories physicists entertain themselves with; they can’t truly know what happened billions of years ago, billions of light-years away, based on the meager data collected by telescopes (which covers of the electromagnetic spectrum). Moreover, in the context of the proposed model, the very ambition implied by the term “cosmology" is at odds with the humility demanded of a physicist, whose entire observable universe could be another physicist’s oven. On the other hand, astronomical observations associated with cosmology, also serve as a laboratory for testing ‘terrestrial’ physical theories, e.g., atomic-, nuclear-, quantum-physics, and this would be particularly true in our case, where the large and the small are so intimately interdependent. When the most compelling cosmological story we can devise, fitting those observations, requires contrived adjustments to terrestrial-physics theories, confidence in those, including GR, should be shaken.
Reluctantly, then, a cosmological model is outlined below. Its purpose at this stage is not to challenge CDM in the usual arena of precision measurements, but to demonstrate how the novel ingredients of the proposed formalism could, perhaps, lead to a full-fledged cosmological model free of the aforementioned flaw.
3.4.1. A Newtonian Cosmological Model
As a warm-up exercise, we wish to solve the system (
57)(
48) for a spherical, uniform, expanding cloud of massive particles originating from the scaling center (without loss of generality). The path of a typical particle is described by
where
a constant vector. It is easily verified that the same homogeneous expanding cloud would appear to an observer fixed to any particle, not just the one at the origin. The mass density of the cloud depends on
a via
, retaining its uniformity at any time and scale if creation/annihilation of matter in scale is uniform across space. The gravitational force acting on a particle is given by
(the uniform vacuum energy is ignored as its contribution to the force can only vanish by symmetry) and (
57) gives a single, particle-independent equation for
a
with
etc.
Two types of solutions for (
73) which are well behaved at all scales should be distinguished: Bounded and unbounded. In the former
is identically zero at
and a
-dependent ‘big-crunch’ time,
. By our previous remarks, at large scale the coarsening terms—those multiplied by
on the r.h.s. of (
73)—dominate the flow and must almost cancel each other or else
a would rapidly blow up with increasing
. The resulting necessary condition for a regular
on all scales is a
-dependent o.d.e. in time, which is simply the time derivative of the (first) Friedmann equation for non-relativistic matter
The
k above disappears as a result of this derivative, meaning that it resurfaces as a second integration constant of any magnitude—not just
. Denoting
bounded solutions in which mass is conserved in time are therefore described by some flow in
parameter space for which
shrinks to zero for
. For example, as
k in (
74) plays the role of minus twice the total energy of the explosion per unit mass, for a scale independent
,
monotonically increases with increasing
.
Given a solution of (
74) at large enough
one can then integrate (
73) in its stable, small
direction, where the scaling piece becomes important, but due to the
constraint, some parts of a solution remain deep in their coarsening regime. The same is true for unbounded solutions, but in this case there is no
to start from, rendering the task of finding solutions more difficult; instead of b.c.
for
, we have
for some initial time,
, and the large-
t asymptotic
for some
. Note the consistency with
for some constant
C, which is an exact solution for the
-free (
73) (and its only solution not wildly diverging in magnitude at large
t). One exception to the hardness of the open-solution (applicable also to closed solutions) is a scaling solution,
, where
is an exact solution of (
74) with
and
. Note that the asymptotic b.c. is automatically satisfied for
. Another is the scale invariant solution of (
73), integrated backwards from
to
, implicitly defining
(integrating forward from
leads to nonphysical solutions).
The Newtonian-cloud model, while mostly pedagogical, nonetheless captures a way—perhaps the only way—cosmology is to be viewed within the proposed framework: It does not pertain to the Universe but rather to a universe—an expanding cloud as perceived by a dwarf amidst it. A relative giant, slicing the cloud’s orbit at a much larger , might classify the corresponding section as, e.g., the expanding phase of a Cepheid/red-giant, or a runaway supernova. An even mightier giant may see a decaying radioactive atom. Of course, matter must disappear in such flow to larger and larger scales—a phenomenon already encountered in the linear case which is further discussed below. The rate (in scale) at which this takes place, in the above models, must be compatible with our analysis of galaxies, where mass was assumed conserved in scale. This would be true for small enough global rate, or if around our native scale, mass annihilation takes place primarily outside galaxies (commencing in a galaxy only after scale flow has compressed it to an object currently not identified as a galaxy).
Suppose now for concreteness that a giant’s section is an expanding star. The dwarf’s entire observable universe would in this case correspond to a small sphere, non-concentrically cut from the star. The hot thermal radiation inside that sphere at
, after flowing with (
16) to
, would be much cooler, much less intense, and much more uniform, except for a small dipole term pointing towards the star’s center, approximately proportional to the star’s temperature gradient at the sphere, multiplied by the sphere’s diameter. Similarly for the matter distribution at
, only in this case the distribution of accumulated matter created during the flow is expected to decrease in uniformity if new matter is created close to existing matter. Thus the distribution of matter at
is proportional to the density at
only when smoothed over a large enough ball, whose radius coresponds to a distance at
much larger than the scale of density fluctuations. This would elegantly explain the so-called dipole problem [
9]—the near perfect alignment of the CMB dipole with the dipole deduced from matter distribution, but with over
discrepancy in magnitude; Indeed, the density and temperature inside a star typically have co-linear, inward-pointing gradients, but which differ in magnitude. Note that a uniform cloud ansatz is inconsistent with the existence of such a dipole discrepancy and should therefore be taken as a convenient approximation only, rendering the entire program of precision cosmology futile. The horizon problem of pre inflation cosmology is also trivially explained away by such orbit view of the CMB. Similarly, the tiny but well-resolved deviations from an isotropic CMB (after correcting for the dipole term) might be due to acoustic waves inside the star.
Returning to the scale-flow of interpolating between ‘a universe’ and a star, and recalling that stands for a spacetime phenomenon as represented by a physicist of native scale s, a natural question to ask is: What would this physicist’s lab notes be? A primary anchor facilitating this sort of note-sharing among physicists of different scales is a fixed-point particle, setting both length and mass standard gauges. We can only speculate at this stage what those are, but the fact that the mass of macroscopic matter must be approximately scale invariant—or else rotation curves would not flatten asymptotically—makes atomic nuclei, where most of the mass is concentrated, primary candidates. Note that in the proposed formalism the elementarity of a particle is an ill-defined concept, and the entire program of reductionism must be abandoned. For if zooming into a particle were to ‘reveal its structure’, even a fixed-point would comprise infinitely many copies of itself as part of its attraction basin.
If nuclei approximately retain their size under scale-flow to large , while macroscopic molecular matter shrinks, then some aspects of spacetime physics (at a fixed-scale section) must change. Instinctively, one would attribute the change to a RG flow in parameter space of spacetime theories, e.g., the Yukawa couplings of the Standard Model of particle physics, primarily that of the electron. However, this explanation runs counter to the view advocated in this paper, that (spacetime) sections should always be viewed in the context of their (scale) orbit; If the proposed model is valid, then the whole of spacetime physics is, at best, a useful approximation with a limited scope. Moreover, an RG flow in parameter space cannot fully capture the complexity involved in such a flow, where, e.g., matter could annihilate in scale (subject to charge conservation); ‘electrons’ inside matter, which in our model simply designate the A-field in between nuclei—the same A-field peaking at the location of nuclei—‘merging’ with those nuclei (electron capture?); atomic lattices, whose size is governed by the electronic Bohr radius , might initially scale, but ultimately change structure. At sufficiently large an entire star or even a galaxy would condense into a fixed-point—perhaps a mundane proton, or some more exotic black-hole-like fixed-point which cannot involve a singularity by definition. Finally, we note that, by definition, the self-representation of that scaled physicist slicing at his native scale s, is isomorphic to ours, viz., he reports being made of the same organic molecules as we are made of, which are generically different from those he observes, e.g., in the intergalactic medium. So either actual physicists (as opposed to hypothetical ones, serving as instruments to explain the mathematical flow of ) do not exist in a continuum of native scales, only at those (infinitely many) scales at which hydrogen atoms come in one and the same size; or else they do, in which case we, human astronomers, should start looking around us for odd-looking spectra, which could easily be mistaken for Doppler/gravitational shifts.
3.4.2. Relativistic Cosmology
In order to generalize the Newtonian-cloud universe to relativistic velocities, while retaining the properties of no privileged location and statistical homogeneity, it is convenient to transfer the expansion from the paths of the particles to a maximally symmetric metric—a procedure facilitated by the general covariance of the proposed formalism. Formally, this corresponds to an ‘infinite cloud’ which is a good approximation whenever the size of the cloud and the distance of the observer from its edge are both much greater than
and
. Alternatively, the cosmological principle could be postulated as an axiom. For clarity, the spatially flat (
), maximally symmetric space, with metric
is considered first, for which the only non-vanishing Christoffel symbols are
The gravitational part,
, of the scaling field, appropriate for the description of a universe which is electrically neutral on large enough scales, i.e.,
, is given by solutions of (
33) which, for the metric (
75), reads
However, the generally covariant boundary condition (
34) “far away from matter" is not applicable here. Instead,
is required to be compatible with the (maximal) symmetry of space—its Lie derivative along any Killing field of space must vanish.
The general form of
consistent with the metric (
75) is
Spatial scaling is taken care of by the metric, hence the vanishing . This implies that, in cosmic coordinates the size of a gravitationally bound system whose outmost matter is deep in its scaling regime, e.g. a galaxy with a flat r.c., also scales as a, rather than in Minkowskian coordinates.
Inserting (
78) into (
77) results is a single equation
Importantly, is a solution when a is of either scaling forms, resp. This exposes the fact that, in the generally covariant setting, the scale-direction of giants could be either (equiv. ) or , depending on the coordinate system and its associated solution for the scaling field. As soon becomes apparent, compatibility with the Newtonian model selects a negative —which at any fixed contains two, free, dependent integration constants, referred to below—and a direction of giants.
An important issue which must be addressed before proceeding, concerns the ontological status of the energy-momentum tensor. In GR, sourcing the Einstein tensor is a phenomenological device, equally valid when applied to the hot plasma inside a star, or to the ‘cosmic fluid’. In contrast,
and the scaling field from which
is derived, both enter (
24) as fundamental quantities, on equal footing with
. To make progress, this fundamental status must be relaxed, and the following way seems reasonable: The fundamental scaling field is written
, with
the above, coarse grained gravitational part, and
the field inside matter. The space averages of the fundamental
and
(derived from
) are written at
,
. That such coarse grained pseudo tensor, respecting the symmetries of the coarse grained metric (
75), has the perfect fluid form, can easily be shown.
Plugging
thus defined and (
75) into the metric flow (
24), results in space-space and time-time components given, respectively, by
with
and
p incorporating
and
while the remaining terms are entirely due to
. Another equation which can be extracted from those two, or directly from (
28), in conjunction with (
79), is energy-momentum conservation in time
Only two of the above three equations are independent due to the Bianchi identity and (
79).
Equations (
80) and (
81) can be combined to
Remembering that paths of co-moving masses can be deduced by analytically continuing solutions of (
24),
, and solving (
59) in the resultant metric (which for the metric (
75) gives:
,
with
a constant), we might as well solve (
83) directly for
. In accordance with
one should also change
(or
), for
in (
83) to be the direction of giants. The result is an equation which, for
, is quite similar to (
73)
only with
multiplying the coarsening piece (due to the different scale-flows involved) and a dark energy term resulting from splitting
, such that for
the scaling piece in (
73) is recovered.
With the above modifications in mind, (
81) becomes
which is the first Friedmann equation with an extra term mimicking dark energy. Since (
85) is satisfied (also) at
, reasonably assuming that
is on the order of the current baryonic density based on direct ‘count’,
, most likely a lower bound, and
based on local measurements (validated below), an estimate
km is obtained, hence
, i.e., the
term in (
85) mimics dark energy which is currently positive.
Let us summarize the computational task of finding a solution for the relativistic cosmological model. The single scale-flow equation is for
a, (
84), whose solutions must be positive, not wildly diverging at large
. Equations (
85), and (
79) with
, act as constraints, which for a given
a and
couple
and
at any (fixed-
) section. The Propagation of
a in scale depends on
p which, as in a standard Friedmann model, requires extra physical input regarding the nature of the energy-momentum tensor, e.g., an equation-of-state relating
p and
. Since both
and
p represent some large-volume average of
, removed of
’s ‘dark’ component, the contribution from inside matter (where
), denoted
, can be assumed to be that of non-relativistic (“cold") matter, i.e.,
. Outside matter the
A-field is nearly a vacuum solution of Maxwell’s equations with an associated traceless
contributing
and
to the total
and
p. If we proceed as usual, identifying
with the energy of retarded radiation emitted by matter, observations would then imply
in the current epoch. However,
incorporates also the ZPF which could potentially even outweigh
. The contribution of the ZPF, being an ‘extension’ of matter outside the support of its
, although having a distinct
dependence, is not an independent component. Properly modeling the combined matter-ZPF fluid, e.g. as interacting fluids, or using some exotic equation-of-state, will be attempted elsewhere.
Returning to the fixed-
constraint, (
85) and (
79), we first note that
in the latter describes the motion of a damped, harmonic oscillator with a negative spring coefficient and a force term (whose sign and magnitude depend on
). A general
negative solution has a single local maxima at
(by
). Since at a fixed
, (
79) is second order in time, only one of its two integration constants is fixed by (
85) evaluated at
. The second one can then be used to further tune
in matching the observed, current acceleration, via
Above,
is the
CDM cold matter density estimate based on supernova- and transverse BAO-distance observations, and
is determined by (
28) (evaluated at
). Thus, the two integration constants of (
79) can conspire to result in an illusion of both a positive cosmological constant, and a cold dark matter addition to
(even if
).
Moving to the early universe, or star-phase of the explosion, at
, the
term in (
85) switches sign, and rapidly decreases with decreasing
t, countering the opposite trend in
, dramatically slowing the shrinkage of
, likely eliminating the horizon problem plaguing a generic Friedmann model. The precise outcome of such a battle of divergences depends on the details of a solution, but a natural, physically motivated scenario follows from the fact that,
,
and
is a solution for the system (
79)(
85) when the two constants are chosen so that the r.h.s. of (
85) vanishes. Namely, the growth of
eventually ‘catches-up’ with that of
, meaning that there is no big-bang in the remote past, just a static universe/star. During that epoch, a pertubative analysis of
,
p,
and
can be performed. A notable departure from standard such analysis is the appearance of ‘vacuum waves’, perturbative solutions of (
36), with associated
masquerading as dark matter of some sort.
Relating cosmological observations to
entails extra steps which are different in the proposed formalism, therefore expected to yield different relations. Remarkably, this isn’t so in most cases. Consider, e.g., the redshift. To calculate the redshift of a distant, comoving object at
, two adjacent, time-ordered points along its worldline are to be matched with similar two points for earth at
. The matching is done by finding two solutions of (
59) which are well behaved on all scales, satisfying the light-cone condition, connecting the corresponding points at
. For the metric (
75) and scaling field (
78), the equation for
(denoting
) and
of each path becomes
subject to the light-cone condition. The two adjacent solutions at
, indexed by
(earlier) and
, trace trajectories
,
and
,
, and the redshift is calculate from the equality
as
. Now, on the two, non overlapping,
parts of their supports,
(almost) satisfy the light-cone condition which, for the highly symmetric metric (
75) implies
. Assuming that
a changes very little over
, the difference between the two integrals in (
88) coming from those end parts is
which would give the standard expression for the redshift in terms of
a. However, in addition there is also an
contribution from the overlap
(rounding the boundary points for clarity which is legitimate to leading order), as the two solutions for (
86)(
87) see slightly different Hubble parameters, on the order of
, and a slightly different scaling field in (
87),
. Nonetheless, since
vanishes at its endpoints, contribution (
90) also vanishes. Note that only the light-condition entered the above analysis, rather than the explicit, conjectured (
86) and (
87). Similarly for the angular diameter distance and the luminosity distance, the latter further requiring exact conservation of
, which is true also in our model.
Finally, the flatness problem of pre-inflation cosmology is elegantly dismissed as follows. First, generalizing the relativistic model to a curved-space FLRW metric is straightforward, and the Friedmann equation (
85) receives a
addition to its r.h.s. Denoting the ratio between its
k and
r.h.s. terms,
, the flatness problem can be stated as the “unrealistic" fine-tuning of
to near mathematical zero at early times, needed to bring its current, observed value to zero within measurement uncertainties. For example, if
, with
encoding the creation/annihilation of matter in scale flow, then
which, at a fixed
, grows by many orders of magnitude over the history of the universe. However, in our formalism the universe is not a machine, propagating in time its state at an earlier time, as previously explained in the context of Bell’s theorem; Friedmann’s equation (
85) enters the relativistic model as a constraint, not an evolution rule, and a cosmological solution is just what emerges out of the set of all constraints. Moreover, even when seen as an evolution rule, (
85) may lead to the following counter argument: At a fixed time, a reasonable
interpolating between a star and a ‘universe’ would counter the growth of
a in the
direction (i.e., the scale-rate of density growth due to matter creation is greater than the third root of its geometric depletion rate). Thus, unless
is fantastically large close to the scale,
, at which a giant’s section corresponds to a star—and why should it be?—
to within measurement uncertainties is a perfectly “realistic".