1. Introduction
The sensitivity of physical processes to the values of fundamental constants, such as the fine-structure constant and the proton-to-electron mass ratio —where and are the proton and electron masses, respectively—has been the subject of extensive theoretical and observational investigation. Variations in these constants impact not only particle physics but also astrophysical processes, the formation of complex structures, and the potential habitability of the Universe.
Comprehensive reviews [
1,
2] provide detailed accounts of how changes in
or
modify atomic energy levels, nuclear reaction rates, and the epochs of recombination and primordial nucleosynthesis. For instance, even a modest increase in
accelerates atomic binding, shifting recombination to higher redshifts and potentially suppressing structure formation by altering the acoustic scale. Conversely, a lower
delays recombination and weakens atomic binding, with cascading effects on the cosmic microwave background (CMB) and baryon acoustic oscillations (BAOs). Anthropic arguments have long explored the constraints imposed by these constants (e.g., [
3,
4,
5]). Other aspects have been investigated in, e.g., [
6,
7,
8,
9,
10].
These and other explorations of how fundamental constants impact astrophysics and cosmology motivate us to question their possible—if sometimes indirect or hidden—role in shaping the velocity fields associated with typical astrophysical systems. Remarkably, velocities observed in such systems never exceed a threshold of a few thousand km/s. For example, the Earth’s velocity with respect to the Hubble frame is ≲400 km/s. The escape velocities from Venus, Earth, Mars, Saturn, and the Sun are approximately 6, 11, 5, 35, and 620 km/s, respectively. The solar system velocity relative to the Galactic center is 220–240 km/s, and typical infall velocities of merging clusters and subclusters have never been observed to exceed a few thousand km/s. The corresponding gravitational wells are characterized by , where is the gravitational potential in speed of light-squared, , units ( is the “compactness” parameter, where and R are its Schwarzschild radius and actual size, respectively) and is the velocity in speed of light units.
This fact is somewhat puzzling, as a much more probable state would be a Universe teeming with black holes (BHs) and neutron stars (NSs), with
and typical velocities approaching the speed of light. This is so simply because the next-to-highest entropy state of a gravitating system is when it forms a BH that eventually evaporates by emitting maximum entropy blackbody radiation. Such a Universe would be thermodynamically closer to the final heat death state, in which all BHs are evaporated via Hawking radiation and the far-future asymptotic de Sitter space is filled with blackbody radiation characterized by the corresponding temperature
eV [
11]. Regardless of its initial state, the penultimate state of the Universe is expected to be populated by BHs and NSs. On small scales, such a Universe would be much less homogeneous than the one we actually observe; higher-entropy gravitational systems are simply more clustered. The puzzle, then, is why our Universe is so regular and homogeneous (
). A key insight developed in this work is that, for entire, virialized (or supported by degeneracy pressure), and naturally formed astrophysical systems—such as stars, planets, and galaxies—the dimensionless
is fundamentally set by microphysical parameters (such as
,
, and other dimensionless ratios), and is independent of the value of the gravitational constant,
G; given the fact that
and
, the Universe is simply not sufficiently old for having all the potentially available nuclear fuel exhausted by now to allow for stars to collapse, the atomic scale that determines the mass density of rocky planets is much larger than the nuclear scale, etc. In other words, non-gravitational interactions are sufficiently effective in inhibiting the Universe from reaching its ultimate fate, even
14 Gyrs after the big bang. The intensive nature of
does not apply to arbitrary subsystems (for example, a fragment cut from a planet), but holds for systems that have reached equilibrium through self-gravity and the relevant microphysical processes. Although the dependence of astrophysical masses and radii on fundamental constants is well explored, the general
G-independence of
for virialized or degeneracy pressure supported systems is not explicitly discussed in the literature, and we aim to highlight and clarify this point throughout the paper.
On a different front, and seemingly a sharp digression, it is worth noting that viewing the Planck mass,
, and gravitational masses,
, through the lens of extensive and intensive properties—contrasted with the Higgs mass,
, and inertial masses,
—could potentially shed new light on the longstanding mass hierarchy problem, e.g., [
12,
13,
14,
15], spanning ∼17 orders of magnitude, as well as the cosmological constant problem, e.g., [
16,
17,
18,
19], of ∼122 orders of magnitude. From this perspective,
and
may be regarded as extensive quantities, while the ratio
remains intensive. Such a perspective leaves the standard gravitational framework intact while potentially offering fresh insights into why these fine-tuning ’problems’ arise in the first place.
This paper is organized as follows. In
Section 2, we summarize a few basic results from the literature on the typical masses, radii, and temperatures of certain virialized or degeneracy-pressure-supported astrophysical systems. In
Section 3 we discuss the relation between extensive formulation of gravity and local scale invariance and how the latter emerges when intensive microscopic description is generalized to extensive formulation of gravity. In
Section 4 we discuss possible implications of extensive Planck mass and we summarize in
Section 5. Throughout, we adopt natural units where the Boltzmann constant,
, Planck constant,
ℏ, and the speed of light,
c, are all set to unity. Our estimates are only crude order-of-magnitude calculations, but this is sufficient for our purposes.
2. Dimensional Analysis Considerations
Dimensional analysis is a powerful technique that generally allows for order-of-magnitude estimates of key quantities, encapsulating the main physical ingredients in a transparent fashion without recourse to differential equations (e.g., [
3,
20,
21,
22,
23,
24,
25,
26,
27]). It captures the qualitative physics and underlying microphysics, leaving the details and precision to the full theory.
In the following, we start with the basic equations to correctly put the context and for future reference but quickly move on to dimensional considerations. The classical equations for gravitating systems include the Poisson equation for the gravitational potential,
,
and the continuity and Euler equations, respectively
supplemented by the equation of state (EOS)
The latter relates the pressure,
P, to the mass density,
, and possibly other parameters denoted by
. Here,
is the velocity field. In realistic systems,
P is provided primarily by non-gravitational interactions or otherwise by quantum degeneracy. However, at the simple-minded level of discussion adopted here, we model the astrophysical system as purely gravitational although the mass density values we adopt (as will be seen in the following) certainly rely on the existence of non-gravitational interactions. To support these over-idealized spherically symmetric, constant density and static systems against their own gravitational pull without recourse to non-gravitationally produced pressure, we are automatically led to model our spherically symmetric static systems as de Sitter solutions of the Einstein field equations in
Section 3.2.
In the following,
is either the (electromagnetic) binding energy per atom, e.g., in the case of planets, nuclear burning stars, galaxies, etc., or the quantum degeneracy pressure in the case of WDs or NSs. The mass density and compactness of an astrophysical system of mass
M and radius
R are given by
where the electromagnetic binding energy is consistent with the virial temperature of the system,
, and
is the Schwarzschild radius (strictly, it would be the corresponding de Sitter scale in our approximation but, given that the current analysis is only an order of magnitude approximation, the distinction is of no significant relevance). It should be stressed that, whereas
is a local property of the astrophysical system, the gravitational potential is not, and
expresses values at the boundaries of entire, naturally formed, systems, with typical size
R. For such systems,
, and it follows that
, provided that
is indeed determined exclusively by microphysics. In this case,
, and so does
—the number of protons that compose the system (assuming negligible binding energy). However, cutting a piece or fragment of size
from the system, e.g., a rock from a planet, then, whereas
is unchanged,
is clearly not
G-independent for an arbitrarily chosen
. It addition, the exterior solution for
is
, and so the maximum value of the gravitational potential at
is
G-independent. The unique ‘status’ of the system boundary that allows for the decoupling of
G from
at this surface is physically clear; typical velocities at the system boundary must not exceed the escape velocity. The latter is bounded by the thermal velocity, which is in turn set by microphysics.
In case that the compactness is determined by the binding energy, then since
, the requirement that the escape velocity,
, is larger than the thermal velocity,
, and that the rms of the latter is determined by the temperature and the mass of the heaviest stable particle, the proton, via
, we obtain that typically
. If both
and
on the right-hand sides of Equations (5) and (6) are determined by the relevant microphysics of the system in question, then the system of Equations (5) and (6) can be solved for
R, resulting in
where
. It then follows that
. Crucially, Equation (
6),
, implies that the compactness,
—not only the density,
—is completely decoupled from
G. The latter only affects the typical values of
M and
R. As pointed out above, assuming
and
are fully determined by microphysics and are independent of
G, it immediately follows that
and
. Exceptions are galaxies (and potentially galaxy clusters as well), whose typical
does depend on
G, and so
M and
R have different
G-dependencies. Nevertheless, since even in these systems
T is independent of
G, then
is independent of
G, and the
G-independence of the corresponding escape/virial velocities, maximum level of gravitational lensing, and gravitational redshift automatically follows.
The temperature,
T, is a property of the system. For example, the temperature of planets cannot exceed the Rydberg energy (in fact, it must be much lower; see [
21]), the temperature of fuel burning stars cannot be smaller than
1 keV, otherwise the Coulomb barrier cannot be overcome to allow for nuclear burning, etc. The idea is that the temperature of virialized systems sets a lower limit on the escape velocity, and in the case where the thermal velocity exceeds the escape velocity the system quickly disperses and disintegrates. However, the escape velocity sets the gravitational potential, i.e., the ratio
. If in addition
is known from the microphysics of the system in question, then, as argued above, these two constraints can be solved for the
characteristic M and
R. It is perhaps worth mentioning that since
, then the highest possible temperatures of naturally formed astrophysical systems cannot exceed
K, otherwise
reaches unity, typical velocities are close to the speed of light, and the system quickly disperses. For comparison, typical temperatures of the intracluster plasma in galaxy clusters fall in the range
K, accretion disks around BHs and NSs are usually not hotter than
K, the temperatures in cores of newly formed NSs are typically
K, etc.
Two immediate consequences of all this are as follows. First, ‘line broadening’ around a frequency , caused by thermal motion of the emitters in a gravitationally bound system—essentially a redshift smearing effec—is maximized at the boundaries of astrophysical systems in a G-independent fashion. Second, although gravitational lensing is sourced by gravity, it is now trivial to see that the maximum deflection angle is independent of G. This simply follows from the fact that the geodesic equations depend only on , but the maximum absolute value of the latter is independent of G, and it then readily follows that the maximal deflection angle is independent of G. The latter could be a billion times larger or smaller than its measured value, yet the maximum value of the deflection angle would not change for it is set by G-independent microphysics. More formally, the deflection angle, , is determined by the transverse impulse, i.e., the ratio of the transverse photon momentum change to its momentum, , where is the gradient transversal to the line of sight. Now, since is typically independent of G at its maximum in virialized and degeneracy-pressure-supported astrophysical systems, then whatever G-dependence length scales might have, the combination remains independent of G at its maximum.
In the following we inspect a few examples more closely. Our first example is rocky habitable planets. Following the arguments of [
21], the temperature of habitable planets cannot exceed a certain threshold, specifically some fraction
of the Rydberg energy,
, so as not to disrupt the chemistry required for sustaining living organisms, i.e.,
, where
. All this implies, e.g., [
21], that
. In other words, since
eV and
GeV, then
. For planet Earth, with radius
km, it implies that the corresponding Schwarzschild radius is
cm. The typical energy density of planets is given by
, where
is the Bohr radius, i.e., each Bohr volume is populated by approximately a single atom. Together, these simple atomic physics-based estimates of
T and
uniquely determine
R and
M, as described in Equations (5) and (6). In the case of planets in general, we require
because the atomic structure of matter has to be preserved, albeit the chemistry of living organisms is no longer a condition that needs to be satisfied. Thus, we require
and so
in this case, and
R and
M are correspondingly larger than in the habitable planets case by factors of
and
, respectively.
Another example is nuclear-burning stars. Based on very general nuclear kinematic arguments a typical stellar temperature
keV is deduced in [
22]. It then immediately follows that
. Here again, it is the weakness of the electromagnetic interaction that allows penetration through the Coulomb barrier and the ensuing nuclear fusion to take place. Had
been significantly larger, nuclear reactions would be impossible, resulting in matter collapsing under its own gravity into a BH with
.
In [
23], the estimate of galactic-scale systems required that, for a gas cloud to fragment and form stars, the typical temperature has to drop below the Rydberg temperature. This by itself determines
. Another constraint used in [
23] was that the dynamical gravitational time is comparable to the (Thomson) cooling time. Together, these constraints resulted in a fair estimate of
M and
R. However, this early analysis ignored the role of cold dark matter (CDM). It was revised later in [
24] to account for CDM. As expected, the conclusion that
is shallow is robust against this improved analysis.
Molecular clouds do not seem to have been discussed in the literature in the context of typical dimensional analysis-based scale estimates. Very general dimensional arguments imply that the vibrational and rotational molecular energy levels are given by and . Considering virialized molecular clouds, the corresponding compactness parameters, , are then and , respectively, once again demonstrating the shallowness of . These estimates are general and decoupled from any assumptions pertaining to the corresponding M and R.
White dwarfs (WDs) and NSs are held against their own gravity by degeneracy pressure. In the case of WDs, the degeneracy is that of electrons with typical spacing of order of the Compton wavelength of electrons, naively , and the corresponding . In the case of NSs, the effective temperature is , where is the pion mass, and so , where is the proton-to-pion mass ratio. Thus, only when gravity is opposed by the strong interaction, or by no opposing force at all, is and the corresponding escape velocities become relativistic.
These results, along with others based on references [
20,
21,
22,
23,
24] for various virialized and degeneracy-pressure-supported astrophysical systems, are compiled and summarized in
Table 1. The table presents typical masses, radii, number densities, and compactness in terms of the fine-structure constant,
, the proton-to-electron mass ratio,
, the proton-to-pion mass ratio,
, the relative electromagnetic-to-gravitational interaction strength,
(where
), and a dimensionless parameter,
.
It is worth noting that, while the primary focus of the present work is the rightmost column showing the compactness parameter for a few typical astrophysical systems, the information in the other columns illustrates the valuable insights gained from basic dimensional analysis. For example, planets are less massive than their host stars fundamentally because , NSs are less massive than their progenitor stars because , and the observed Sun-to-Jupiter mass ratio, , is simply determined by . Likewise, the star-to-habitable planet (such as Earth) mass ratio is given by . While these observed ratios are typically explained by the details of various astrophysical processes, these processes are ultimately governed by dimensionless ratios of fundamental constants of nature, as demonstrated here. It is also important to realize the fact that is not a result of the relative weakness of gravity in comparison with the other interactions gauged by, e.g., , but rather a result of the weakness of the latter in absolute terms, i.e., the fact that , , etc.
As discussed above and further demonstrated in
Table 1, gravity, as encapsulated by the gravitational constant,
G, primarily determines the overall masses, radii, and particle numbers of gravitationally bound systems, setting the scale for the existence of macroscopic astrophysical objects. However, once such systems form and reach virial equilibrium, the maximum depth of their gravitational wells, as characterized by the dimensionless compactness,
, is dictated by microphysical parameters and remains fundamentally independent of the value of
G. Whereas the peak values of the rms velocities in and around these systems is
G-independent, these velocities do depend on
G well within or outside the system boundary. In this sense,
G enables the presence of large-scale structures, but does not directly govern the maximum characteristic velocities or compactness of these virialized systems.
In hindsight, the
G-independence of
becomes even more transparent when derived directly from Equation (
6), since
G does not enter the derivation; both
and
are evidently independent of
G. The results of this direct approach can be contrasted with the estimates in the rightmost column of
Table 1 for comparison and verification. For instance, the typical binding energy of rocky planets is the Rydberg energy,
, as they are composed of atomic matter. Consequently,
in this case. Compactness parameters,
, are shown for reference in
Table 2 for planets in our solar system. Similarly, gas cloud fragmentation in star-forming galaxies depends on the gas temperature, with the Rydberg energy serving as a threshold [
23], implying that typical galactic compactness is comparable to that of planets,
. For habitable planets, the binding energy is further reduced by a factor
, as discussed in [
21]. In the case of WDs, the relativistic electron degeneracy pressure balancing gravity corresponds to an energy scale
, where
is the typical electron momentum uncertainty consistent with Heisenberg’s principle, and
d is the electron Compton wavelength,
. Thus,
. For NSs, the non-relativistic degeneracy pressure yields
, with
, resulting in
, which lies within the range
, shown in
Table 1.
Since
at the system boundary can be inferred from fundamental physics directly, with no reference to estimates of
M and
R, i.e., with no reference to
G, the possibility of reformulating gravitation in a
G-independent fashion – for that purpose only—could be useful. Consider the classical gravitational attraction between two masses,
and
. The mutual attraction force
can be cast in the form
, where
is the relative distance in Planck units, and
is the Planck length. This
G-independent description, where distances are measured in Planck units rather than standard meters, inches, etc., is especially suitable for a
G-free description of
. We demonstrate in
Section 3.3 below how such a reformulation of GR can be achieved.
4. Implications of Extensive Planck Mass
Having experimented with the intensive nature of the compactness parameters of typical astrophysical systems, , and with the emergence of the WI version of GR and the emergence of a macroscopic description starting from a microscopic G-independent description, we are conveniently positioned to explore the implications of the possibility that the experimentally inferred Planck mass is actually an extensive property of astrophysical systems.
In purely Cavendish-type tests, only the combination
is being measured. Notably, under arbitrary rescaling
and
, where
k is an arbitrary, universal rescaling function, the measured
remains invariant. The conventional choice,
, corresponds to the imposition that gravitational mass that appears in the universal law of gravitation is identified with the inertial mass. It is then possible to nominally define the Planck mass by
. However, this nominal value is purely convention-dependent. Notably, this ‘scale invariance’ of
is actually naturally built in the WI version of GR that is discussed in
Section 3.3, where the modulus of the scalar field,
, regulates both the Planck mass and gravitational masses in a way that leaves the coupling strength
invariant.
As discussed above, a subtle but significant aspect of gravitational physics is the fashion in which the gravitational constant, G, enters observable phenomena. In many gravitational systems—particularly those that are virialized, such as stars, planets, and galaxies—upper bounds on quantities accessible to observation (e.g., gravitational redshift, line broadening, lensing deflection angles, escape velocities, and virial velocities) are determined by microphysics with no reference to G. It is only in the case of non-virialized systems, or subsystems of otherwise virialized (or degeneracy pressure-supported) systems, or in the direct interaction between two effectively pointlike masses (e.g., the Earth–Sun system), that G appears explicitly in observable quantities, as in Newton’s gravitational law. In these cases, the force or acceleration is directly proportional to G and can be measured unambiguously.
This pattern suggests that the role of
G is, in a sense, emergent; it is hidden in configurations where gravitation is balanced by other non-gravitational interactions and revealed only when comparing/measuring gravitational effects to/by non-gravitational standards or when considering pointlike interactions. This conclusion is ingrained and manifested throughout the analysis in
Section 3. In this context,
G serves as a conversion factor between energy density that is set by microphysics and the macroscopic spacetime curvature, but its absolute value is often inaccessible in the absence of a fixed external non-gravitational scale. This observation highlights a profound feature of gravitational physics, with possible implications for our understanding of the fundamental vs. emergent nature of gravity.
It should be stressed that the absolute value of G is primarily determined by laboratory experiments on Earth, such as Cavendish-type measurements, and even then it is the measurement of gravitational interaction by comparison with non-gravitational forces that allows the empirical determination of G. Solar system observations—including lunar laser ranging, planetary orbit tracking, and spacecraft Doppler data—measure the product (where M is the mass of a celestial body) and cannot independently determine G without knowing M. Therefore, these methods mainly constrain temporal or spatial variations of G relative to the Earth-based value, rather than empirically inferring its absolute magnitude.
The following example illustrates one way in which
G might emerge. As discussed below Equation (
6)—and shown with specific cases in
Table 1—systems whose characteristic density,
, is fixed by microphysics (and does not involve
G) exhibit scalings
and
. By contrast, systems (like galaxies) whose formation process depends on
G have
, leading instead to
and
as required by the
G-independence of
.
Here we focus on the more generic case where
truly does not depend on
G, and in particular on habitable planets. Our objective is to explain—at the dimensional analysis level—why terrestrial Cavendish-type experiments measure an effective Planck mass
GeV rather than some other value. From
Table 1, one finds that the typical radius of such a planet is
, where
is the Bohr radius,
the fine-structure constant, and
a parameter of order
, as was already defined in
Section 2 but we reiterate for convenience. Since atoms in a rocky planet are tightly packed, we also have
, where
N is the number of atoms. Combining these two estimates to eliminate
R, we arrive at
.
Conventionally, we would choose
as defined at the beginning of this section, i.e., we set the gravitational mass equal to the inertial mass thereby fixing
G, and interpret this relation as saying that the number of atoms,
N, in a habitable planet is determined by
,
,
, and
. Indeed, a similar approach has been adopted in the context of estimating the number of protons in a typical nuclear burning star, e.g., [
20,
31,
32]. However, adopting the viewpoint that
is merely an emergent conversion factor, and in addition noting that empirically
is of order of the Higgs mass,
, this logic can be reversed, and we are led to tentatively conjecture instead that
i.e., that
is determined by
N and not the other way around—the larger
N is the weaker is gravity, and the larger is
. From this perspective, the Earth-based laboratory determination of a small
G (equivalently a large
) simply reflects the fact that typical habitable planets like Earth contain
atoms, producing a
-order-of-magnitude hierarchy
. Following this logic and repeating Cavendish-like experiments on, e.g., the moon or Mars (while fixing
) will result in different inferred values for
G (i.e., values different from that inferred on Earth) if our assumption that
G is an emergent value following Equation (
13) rather than a universal constant is correct. This is a falsifiable prediction of our proposal. It should be also emphasized that the proposed Planck-to-Higgs mass relation, Equation (
13), is not rigorously derived from any first princples and it may well be an empirical coincidence that it empirically fits Earth-based measurements. However, the scaling
is more robust as was just discussed above Equation (
13), and the proportionality factor can be some other non-Higgs scale. Nevertheless, it would be a remarkable coincidence if the latter just happens to nearly coincide with
.
Even the number of particles in Earth is determined by fundamental principles to be
. Fitting the order-of-magnitude Earth compactness
(as shown in
Table 1) to its measured compactness, we obtain that
. Combining the proposed scaling
with
, we than conclude that
. Assuming
and the measured proton mass, we remarkably obtain the order-of-magnitude estimate
GeV. Crucially, within each system, the ratio
is independent of
N because both gravitational mass and Planck mass scale identically with
. This ensures gravitational interactions between any two systems with different particle numbers remain universal and consistent with the equivalence principle, as the gravitational coupling between two systems with masses
and
depends only on these ratios as
.
Thinking in terms of current terrestrial experimental standards and precision of
G-inference could be disheartening. Indeed, conducting a Cavendish-like experiment on, e.g., the moon or Mars is a daunting task that presents both intriguing scientific opportunities and significant challenges. Although technically demanding, successfully conducting this experiment could provide significant insights into the nature of gravity, making it a scientifically valuable endeavor worth pursuing despite the obstacles. This is especially true given that falsifying either the universality of
or its proposed
dependence on, e.g., the moon is not as demanding as a full-fledged precision inference of
G with current standards of precision measurements. Specifically, according to Equation (
13), the fractional relative divergence of
inferred in two systems that contain
and
particles, respectively, is expected to be
. With the moon-to-Earth mass ratio of ∼0.012 and assuming small binding energies in both systems we obtain
, and consequently a very modest ∼10 percent precision is already more than sufficient to falsify the proposed scaling.
Even on Earth,
G can vary from place to place on its surface, because the surface of planets is shaped not only by gravity but also by the electromagnetic interaction as well, i.e., by the strength of matter as measured by its bulk modulus (compressibility). On planets, this is typically of order
, e.g., [
22]. The estimated mass of Mount Everest, approximated as a cone with a base radius of about 5 km and height 8.8 km, and assuming an average Earth density of 5 gr/cm
3, is ∼10
18 gr. Comparing this with Earth mass
gr and assuming similar mass densities for both, we find
, the fractional Planck mass difference between sea level and the summit of mount Everest is then
, which is smaller than one part in ten billion. This is a much smaller difference than current precision in
G measurements. The latter is not much better than one part in
. Thus, whereas a moon-based Cavendish-like experiment can put Equation (
13) to test, near-future Earth-based falsification of this conjecture seems to be not feasible.
One immediate consequence of the scaling
is that the Higgs-to-Planck energy-density ratio becomes
On cosmological scales where
is dominated by CMB photons and primordial neutrinos this ratio is
. This is tentatively close to the ∼122 orders-of-magnitude fine tuning of the observed
, i.e., the energy density associated with the cosmological constant, and the naive expectation for the zero-point vacuum energy density,
[
16]. From the present work perspective, and as we explicitly saw in
Section 3, the cosmological description of a cosmic evolution dominated by
, i.e., a vacuum energy-like cosmological constant, is in the microscopic, intensive, frame that corresponds to
. Compared with the extensive frame description with
, the ∼122 orders of magnitude gap between observations and naive theoretical expectations can be almost fully bridged by realizing that the apparent mismatch results from comparing intensive with extensive quantities in a very complex system that contains
particles.
Although the extensive vs. intensive perspective requires a markedly unusual treatment of G, which admittedly diverges even from other emergent-gravity perspectives, it offers a new perspective—by virtue of its N-dependence—on both the Planck–Higgs mass hierarchy and the ‘unexpectedly’ small cosmological constant.
At first glance, it may seem that we have simply traded the Higgs–Planck mass hierarchy and the cosmological constant problems for the question: why is our Universe so vast in complexity, i.e., why do habitable planets, nuclear-burning stars, and the observable Universe contain , and particles, respectively? However, our interpretation offers a unified, quantitative scaling of G with system complexity N, thereby bringing both hierarchies under one umbrella. The ultimate (qualitative, if not quantitative) resolution of “why so large complexity?” may ultimately—perhaps unsurprisingly—be anthropic; only a sufficiently complex cosmos can give rise to complex observers.
5. Summary
Typical astrophysical velocities are almost always non-relativistic, and the gravitational wells
characterizing these astrophysical systems are correspondingly shallow, i.e.,
, where
and
R are their typical Schwarzschild radius and actual size, respectively. According to the generalized Second Law of thermodynamics, this implies that the entropy budget of our Universe is much smaller than the holographic entropy bound,
[
11].
This can be neatly explained by the fact that astrophysical systems are either regulated by cooling processes (governed by the electromagnetic interaction) or are required not to allow certain electromagnetic processes to occur. In other words, these systems must remain below specific temperature thresholds, otherwise their virial velocities would exceed their escape velocities and the systems would rapidly disperse. Typical temperatures are set by atomic or molecular physics, with , so that virialized objects are typically characterized by , where and is positive.
Had and been of order unity, the entire Universe would be teeming with BHs and NSs, with typical velocities approaching the speed of light, shortly after the Big Bang—regardless of its initial entropy state. So why is the Universe not predominantly populated by BHs and NSs? Simply because and .
Whereas the reasoning laid out in this work is straightforward, the somewhat surprising conclusion—that, although the existence of astrophysical systems depends on G and specifically on its empirical value, large-scale velocities at the system boundary are determined solely by microphysics, essentially by the electromagnetic interaction rather than by gravitation—is not sufficiently appreciated in the literature. One objective of this work was to emphasize this rather surprising fact. Although order-of-magnitude estimates of M, R, and T for various astrophysical systems from first principles are discussed in the literature, the facts that is small due to the weakness of the electromagnetic interaction and the small electron-to-proton mass ratio, and is independent of G, seem to be overlooked, or at least not sufficiently acknowledged. In particular, we stress that this G-independence of applies specifically to entire, virialized, naturally formed astrophysical systems—such as stars, planets, and galaxies—that have formed and equilibrated through self-gravity and the relevant microphysical processes. Arbitrary non-virialized subsystems, like a piece cut from a planet, do not exhibit this property.
In gravitational systems, distinguishing between intensive and extensive quantities offers a coherent framework to tentatively address the mass hierarchy and dark energy problems. Specifically, the inertial mass scale,
, set by particle physics, is a constant of nature, independent of system size or particle number,
N. In contrast, we put forward the proposal that the gravitational mass scale, characterized by the effective Planck mass,
, scales as
(ignoring possible geometric factors—or others—of order unity), a conjecture that could be directly falsified by future moderate-precision moon-based Cavendish-like experiments. For example, Earth’s particle content,
, quantitatively explains the observed hierarchy
measured in Cavendish experiments, linking microscopic particle masses to macroscopic gravitational scales. Extending this to cosmology, vacuum energy density,
, is measured as we saw in
Section 3 as an intensive quantity, i.e.,
. With
, approximately the number of CMB photons in the observable Universe, this scaling naturally suppresses naive particle physics predictions by about 120 orders of magnitude; observing the Universe from within, cosmological measurements probe intensive quantities that reflect the gravitational scaling tied to the Universe particle content, e.g., the energy density rather than total mass of the Universe (much like the energy density of the CMB does not depend on the total number of CMB photons in the Universe and the mass densities of, e.g., planets, WDs and NSs depend exclusively on microphysics). As a consequence, the ‘naturalness’ problems associated with the mass hierarchy and cosmological constant could be rephrased as to why our Universe so complex, i.e., why, e.g., planet Earth and the observable Universe contain
and
particles, respectively. This question is rarely asked in the context of these puzzling problems.