Preprint
Article

Firewalls, Hawking (Tolman) Radiation, and a Tentative Resolution of the Firewall-Mass Problem

Altmetrics

Downloads

199

Views

64

Comments

0

This version is not peer-reviewed

Submitted:

04 March 2024

Posted:

05 March 2024

You are already at the latest version

Alerts
Abstract
It has been theorized that black holes are surrounded by firewalls, although there is not universal agreement concerning this. We first review basic concepts pertaining to Schwarzschild black holes and Hawking radiation. Then we discuss the anticipation of Hawking radiation—albeit from non-black holes—initially by R. C. Tolman and shortly thereafter with P. Ehrenfest. We compare evaporation into a vacuum at absolute zero (0 K) of black holes with that of non-black holes, and show that not only black holes but also non-black holes evaporate within a finite time. The times required for evaporation of black holes and non-black holes are compared. Next, we show that (i) if firewalls exist, they can originate via Hawking radiation at the minimum possible ruler distance (the Planck length) beyond the Schwarzschild horizon, where it has not suffered any gravitational redshift, or, alternatively, suffered maximal gravitational blueshift and (ii) the firewall temperature is on the order of the Planck temperature, independently of the mass and hence also of the Schwarzschild radius of a Schwarzschild black hole. We then explain the exponential nature of the gravitational frequency shift as a function of the gravitational potential. Next, we consider the firewall-mass problem, and provide an at least prima facie tentative resolution thereto based on: (i) the mass of a firewall being canceled by the negative gravitational mass = (negative gravitational energy)/c2 accompanying its formation, (ii) the unchanged observations of a distant observer upon formation of a firewall, and (iii) Birkhoff's Theorem (actually first discovered by Jørg Tofte Jebsen). We then consider one aspect of thermodynamics in gravitational fields, showing that equilibrium relativistic gravitational temperature gradients cannot be exploited to violate the Second Law of Thermodynamics.
Keywords: 
Subject: Physical Sciences  -   Astronomy and Astrophysics

1. Introduction

It has been theorized that black holes are surrounded by firewalls [1–7], although there is not universal agreement concerning this [1–7]. There is a vast literature exploring this topic, of which we have cited only a tiny sample. But the many works cited, and discussed, in our cited Refs. [1–7] could provide at least the beginning of a thorough literature search [1–7].
In Section 2, we review basic concepts pertaining to Schwarzschild black holes and Hawking radiation. In Section 3, we discuss the anticipation of Hawking radiation—albeit from non-black holes—initially by R. C. Tolman and shortly thereafter with P. Ehrenfest, and Tolman’s [8,9] proof, bolstered with Ehrenfest [10], that any gravitator—black hole or non-black hole—must radiate and hence cannot be in thermodynamic equilibrium with a surrounding vacuum at (or at least sufficiently close to) absolute zero ( 0   K ), but must completely evaporate into that vacuum within a finite time. The times required for evaporation of black holes and non-black holes are compared. (See also Garrod [11].) This has been corroborated by recent research [12]. In Section 4, we show that (i) if firewalls exist, they can originate via Hawking radiation at the minimum possible ruler distance [13] (the Planck length [14–16]) beyond the Schwarzschild horizon, where it has not suffered any gravitational redshift [17,18], or, alternatively, suffered maximal gravitational blueshift [17,18] and (ii) the firewall temperature is on the order of the Planck temperature [19], independently of the mass and hence also of the Schwarzschild radius of a Schwarzschild black hole. We emphasize and focus on ruler distance [13] because, unlike other distance measures in General Relativity [13], ruler distance—uniquely! [13]—is actual physical distance: the distance between two points as measured by rulers laid upon the shortest possible spatial path separating them and hence the physical distance separating them [13]. (We employ other distance measures where they are more applicable.) In Section 5, we explain the exponential nature of the gravitational frequency shift as a function of the gravitational potential. In Section 6, we consider the firewall-mass problem [3], and provide an at least prima facie tentative resolution thereto based on: (i) the mass of a firewall being canceled by the negative gravitational mass [17,18] = ( negative gravitational energy ) / c 2 [17,18] accompanying its formation, (ii) the unchanged observations of a distant observer upon formation of a firewall, and (iii) Birkhoff’s Theorem [20–31]—actually first discovered by Jørg Tofte Jebsen [25–28]. We show that the mass of a firewall is exactly counterbalanced by the (negative) gravitational mass-energy accompanying its formation. Perhaps this may complement other lines of reasoning [4] disputing massiveness [3] of firewalls. (There is a caveat [28–31]   1 with respect to Birkhoff’s Theorem [20–31], but it [28–31]   1 is not relevant with respect to our considerations.   1 ) In Section 7, we consider one aspect of thermodynamics in gravitational fields, showing that equilibrium relativistic gravitational temperature gradients cannot be exploited to violate the Second Law of Thermodynamics. A concluding synopsis is provided in Section 8. Auxiliary topics are discussed in the Notes.

2. Review of Basic Concepts Pertaining to Schwarzschild Black Holes and Hawking Radiation

The Schwarzschild metric of a Schwarzschild black hole of mass M and Schwarzschild radius r S = 2 G M / c 2 is [32]
d s 2 = 1 2 G M r c 2 c 2 d t 2 1 2 G M r c 2 1 d r 2 r 2 d θ 2 sin 2 θ d ϕ 2 = 1 r S r c 2 d t 2 1 r S r 1 d r 2 r 2 d θ 2 sin 2 θ d ϕ 2 = c 2 d τ 2 d l 2 r 2 d θ 2 sin 2 θ d ϕ 2 .
Schwarzschild-coordinate radial distance r is not radial ruler distance but is radial area distance and also radial distance from apparent size [13]. By contrast, in accordance with the Euclidean form of the angular term of the Schwarzschild metric [the last term in all three lines of Equation (1)], a spherical shell at r shell has ruler-distance circumference C shell = 2 π r shell and ruler-distance surface area A shell = 4 π r shell 2 [13]. At r r S , lnotr—is radial ruler distance, t is Schwarzschild-coordinate time (proper time measured by a clock at rest at r ) [17,18], and τ is proper time measured by a clock at rest at any given r r S [17,18]. [A clock at rest at  r = r S —if such a clock can exist—must be constructed entirely of photons (and/or other zero-rest-mass particles)!]
Although not necessary for our derivations, it may be helpful, as an aside, to briefly remark on the following three features of the Schwarzschild metric [Equation (1)]: (i) Setting d s 2 = 0 in Equation (1) shows that the physical radial velocity of light V phys , light = d l / d τ = c at all r r S [32]; but, by contrast, the Schwarzschild-coordinate radial velocity of light V coor , light = d r / d t = c 1 r S / r decreases monotonically with decreasing r from c at r to zero at r = r S [32]. (ii) We focus on distance [13], especially on ruler distance [13], and most especially on radial ruler distance [13], beyond the Schwarzschild horizon r S in Schwarzschild spacetime [32]. But it may be interesting to note that, in accordance with the angular term of the Schwarzschild metric [the last term in all three lines of Equation (1)] being of the identical Euclidean form at all  r 0 [32],   2  even within  r S ,   2 where a spherical shell cannot be at rest but must be collapsing, while falling through a given r shell < r S it has ruler-distance circumference C shell = 2 π r shell and ruler-distance surface area A shell = 4 π r shell 2 [32].   2 Even within r S ,   2 where r becomes timelike, r does not become time itself:   2 unlike time itself r still retains these spatial geometrical attributes [32].   2 Time itself has no spatial geometrical attributes. (iii) Because the gravitational field of a Schwarzschild black hole is purely radial, it seems intuitive that this gravitational field’s stretching [33,34]   3 of space from the Euclidean [33,34]   3 is purely in the vertical   3 radial r direction, and not at all in the horizontal angular θ and ϕ directions: thus the identical Euclidean form of the angular term of the Schwarzschild metric [the last term in all three lines of Equation (1)] at all  r 0 [32]. Indeed, more generally, intuition suggests that stretching [33,34]   3 of space from the Euclidean by any gravitational field is purely in the vertical direction [33,34],   3 and not at all in any horizontal direction [33,34].   3 This intuition augments the immediately preceding Item (ii), and is perhaps most clear in relation to Sakharov’s elastic-strain theory of gravity [35–40].   4 , 5 (Thus, might space be the ether [41,42]?   4 , 5 )
We will be concerned only with radial motions of photons in the gravitational fields of Schwarzschild black holes, because only radial motions can result in gravitational frequency shifts. In this regard we will be concerned only with the temporal-radial part of the Schwarzschild metric [Equation (1)], hence ignoring the angular term thereof (the last term in all three lines thereof).
Hawking radiation, the (at least essentially) blackbody radiation from a Schwarzschild black hole, is most typically construed to have the temperature [16,43–49]
T H , r = c 3 8 π G k M = c 3 8 π G k r S c 2 2 G = c 4 π k r S ,
where k is Boltzmann’s constant, M is the mass of the black hole, and r S = 2 G M / c 2 is its Schwarzschild radius [16,43–49]. Note that T H , r given by Equation (2) is the temperature of Hawking radiation at a great distance from a Schwarzschild black hole, i.e., at r , hence after Hawking radiation having suffered the maximum possible gravitational redshift [16–18,43–49]. For all non-primordial black holes, which are all of stellar mass or larger, T H , r is extremely low compared to the current temperature T CBR = 2.725   K [50] of the cosmic background radiation [50]. For sufficiently small primordial black holes [51–61] ( M < c 3 8 π G k T CBR = 4.50 × 10 22 kg r S < c 4 π k T CBR = 6.69 × 10 5 m ), T H , r > T CBR = 2.725   K obtains. But as of this writing, to the best knowledge of the author, no such sufficiently small primordial black holes—indeed, no primordial black holes at all—have been discovered [51–61]. Moreover, while there are rationales according to which primordial black holes might contribute, perhaps significantly, to cold dark matter, there also are both theoretical and observational upper limits on their abundance [51–61] and therefore also on their actual contribution to cold dark matter [51–61]. Hence, while it is possible that they could contribute, perhaps significantly, to cold dark matter, as of this writing, to the best knowledge of the author, it is uncertain whether or not they actually exist [51–61].
Thus far we have considered T H , r . But T H , r at smaller values of r ( r S r < ) has been discussed as well [49]. Closer to r S (at r S r < ) [49] than at r , Hawking radiation has suffered less [17,18,49] gravitational redshift [17,18] and hence has a higher [49] temperature [16,43–49]
T H , r = T H , r 1 r S r 1 / 2 = c 3 8 π G k M 1 r S r 1 / 2 = c 4 π k r S 1 r S r 1 / 2 .

3. Tolman’s Anticipation of Hawking Radiation: All Gravitators—Black Holes and Non-Black Holes—Must Radiate

It is extremely important to note that Equation (3) is a special case of the more general result derived by Tolman [8,9], bolstered with Ehrenfest [10]. The last two terms of Tolman’s Equation (128.6) and the entirety of Tolman’s Equation (129.10) in Ref. [9] read:
T T g t t 1 / 2 = C T T T = C T g t t 1 / 2 ,
where (i) C T is Tolman’s constant in Equations (128.6) and (129.10) of Ref. [9], and (ii) T T is the Tolman gravitational thermodynamic-equilibrium temperature as measured by a local observer, written more explicitly as T T r , θ , ϕ , at a specified location [specified radial area distance and also radial distance from apparent size [13] r and specified direction θ , ϕ ] from the center of mass of a gravitator (spherically symmetrical or otherwise) where the time-time component of the metric has the value g t t r , θ , ϕ . [As per the first term of Equation (128.6) in Ref. [9] (this notation is also employed in Refs [8] and [10]), Tolman (with Ehrenfest in Ref. [10]) sets e ν = g t t , but for uniformity in notation we employ only the g t t symbolization. Tolman [8,9] (with Ehrenfest in Ref. [10]) employs the symbols T 0 instead of T T (with the subscript   0 designating measurement by a local observer: see p. 489 of Ref. [9]), and C instead of C T . (In Tolman’s paper with Ehrenfest [10] const. is employed instead of C.) We employ the subscript   T to refer to Tolman’s quantities. We take T T to be the Tolman gravitational thermodynamic-equilibrium temperature as measured by a local observer (the subscript   0 omitted for brevity).]
We can choose Tolman’s constant C T for a given gravitator to be equal to C T , r = lim r T T r , θ , ϕ g t t r , θ , ϕ 1 / 2 = T T , r , the Tolman gravitational thermodynamic-equilibrium temperature measured by a local observer at radial area distance and also radial distance from apparent size [13] r in any specified direction θ , ϕ from the center of mass of any gravitator (spherically symmetrical or otherwise), because spacetime approaches the Minkowskian and hence g t t , r = lim r g t t r , θ , ϕ = 1 at radial area distance and also radial distance from apparent size [13] r in any specified direction θ , ϕ from the center of mass of any gravitator (spherically symmetrical or otherwise):
C T = T T , r g t t , r 1 / 2 = T T , r × 1 = T T , r .
Hence Equation (4) can be rewritten as:
T T r , θ , ϕ = T T , r g t t r , θ , ϕ 1 / 2 ( in general ) T T r = T T , r g t t r 1 / 2 = T T , r 1 r S r 1 / 2 ( spherical symmetry ) .
Tolman presents this general result not only via the last two terms of Equation (128.6) and the entirety of Equation (129.10) in Ref. [9], but also in Ref. [8] via the second equation in the Abstract and Equations (27), (28), (42), (53), and (54). It is also presented in Ref. [10] with Ehrenfest via Equations (2), (3), and (30), and discussed in some detail in Section 7 thereof. [In the weak-field limit ( M r S c 2 / 2 G r r S = 2 G M / c 2 given spherical symmetry) this result is presented via Equations (8) and (29) and the last (unnumbered) equation in Section 8 of Ref. [8], and also via Equations (128.4), (128.5), and (128.10) in Ref. [9].] Tolman also evaluates the gradient d ln T T r / T T , r / d r at and near Earth’s surface in the last (unnumbered) equation in Section 2 of Ref. [8] and in Equation (128.7) of Ref. [9]. (In Earth’s weak gravitational field, this gradient can of course with at most negligible error be construed as being either with respect to radial ruler distance [13] or with respect to radial area distance and/or radial distance from apparent size [13].) We focus on the last two terms of Equation (128.6) in Ref. [9] and of Equation (3) in Ref. [10], on Equation (129.10) in Ref. [9], and on Equation (30) and Section 7 in Ref. [10]. (See also Garrod [11].)
For Schwarzschild black holes, we should expect that T T = T H . Indeed, letting T T T H , the second line of Equation (6) is identical to Equation (3), taking T T , r = T H , r as per Equation (2). For T T , r of Schwarzschild (spherically-symmetrical, non-rotating) non-black holes, we will give a plausibility argument (albeit not a proof) for a conjecture concerning the value of T T , r .
We define a Schwarzschild non-black hole (of mass M) as a spherically-symmetrical, non-rotating gravitator whose radius r NBH (as per radial area distance/radial distance from apparent size [13]) exceeds the Schwarzschild radius r S = 2 G M / c 2 . (The subscript   NBH may be omitted for brevity when that will not result in confusion.)
The apparent gravitational acceleration G BH r towards a Schwarzschild black hole of mass M BH , as measured by dangling a unit mass m at r S from a higher altitude r > r S (with a massless string), and the limit thereof as r , are [62]:
G BH r = c 4 4 G M BH 1 r S 2 r 2 1 / 2 = G M BH 2 G M BH c 2 2 1 r S 2 r 2 1 / 2 = G M BH r S 2 1 r S 2 r 2 1 / 2 lim r G BH r G BH = c 4 4 G M BH = G M BH r S 2 .
The limiting value of G BH r as r —as per the second line of Equation (7)—is sometimes called the surface gravity of a Schwarzschild black hole [62]. Fortuitously, as if Newtonian theory was adequate for Schwarzschild black holes [62]! Fortuitously despite Newtonian theory taking all distance measures to be equivalent—not distinguishing, for example, between ruler distance [13] on the one hand, and area distance and/or distance from apparent size [13] on the other. Fortuitously because enormous values of G BH r if r is only slightly greater than r S suffer gravitational redshift down to Newtonian values in the limit r . (Note: We employ g to denote components—we focus on the time-time component—of the spacetime metric, G to denote the universal gravitational constant, and G to denote acceleration due to gravity.) For a Schwarzschild (spherically-symmetrical non-rotating) non-black hole of mass M NBH in the weak-field limit ( M r S c 2 / 2 G r r S = 2 G M / c 2 )
G NBH = G M NBH r NBH 2 .
Based on the fortuitous Newtonian-equivalence of the forms of the second line of Equation (7) on the one hand and Equation (8) on the other, and applying Equation (2), we prima facie suggest the following conjecture for T T , r of a Schwarzschild non-black hole of the same mass M as a Schwarzschild black hole ( M NBH = M BH = M ) but with radius (radial area distance and also radial distance from apparent size [13]) from the center to the surface of the Schwarzschild non-black hole) r NBH > r S = 2 G M / c 2 :
T T , r = T H , r G NBH G BH = c 3 8 π G k M × G NBH G BH for all r NBH > r S = T H , r G M r NBH 2 G M r S 2 = T H , r r S r NBH 2 = c 3 8 π G k M r S r NBH 2 weak - field limit : r NBH r S .
We recognize that while the conjecture given by Equation (9) may, at least prima facie, seem plausible, we have not proven it. Nonetheless at least prima facie it seems a reasonable conjecture, especially given that, since
lim r NBH r S T T , r = T H , r = c 3 8 π G k M ,
at least it is consistent with Equation (2). However, irrespective of the validity (or lack thereof) of this conjecture, two paragraphs hence we will prove the extremely important point that T T , r  must be finitely greater than absolute zero ( 0   K ) for all non-black holes [as T H , r is finitely greater than absolute zero ( 0   K ) for all black holes].
In Ref. [8] and in Sections 128 and 129 of Ref. [9], Tolman implies that our Equations (4)–(6) are valid in any static spacetime [63]. Together with Ehrenfest [10] this is also implied in Ref. [10]. But given rotation at constant angular velocity, a time-independent centrifugal potential can be incorporated into the time-independent gravitational potential that obtains in static spacetime, i.e., into that which obtains neglecting the rotation [63]. This time-independent gravitational-centrifugal potential would then of course be a function of θ as well as of r, but at a given θ it can still be expressed as a function of r alone. Thus we can construe Equations (4)–(6) to be valid in any static or stationary spacetime [63]. Hence these results proven by Tolman [8,9], bolstered with Ehrenfest [10], and summarized via our Equations (4)–(6) and the associated discussions, imply that at thermodynamic equilibrium temperature increases downwards   6 in any static or stationary gravitational field (the centrifugal contribution in a stationary field construed as incorporated therein given rotation at constant angular velocity) [63]. But for simplicity and definiteness we focus on the static spacetimes at r r S of Schwarzschild, i.e., spherically-symmetrical non-rotating (black hole and non-black-hole) gravitators.
Even more importantly, Tolman [8,9], bolstered with Ehrenfest [10], furthermore implies more than that, as summarized via our Equations (4)–(6): it is furthermore implied [8–10] that T T , r  must be finitely higher than absolute zero ( 0   K ) for allnon-black holes, as T H , r isfinitely higher than absolute zero ( 0   K ) for all black holes. As per Equations (4)–(6) [most explicitly as per Equation (6)], incorrectly assuming that T T , r = 0   K incorrectly implies that T T r , θ , ϕ = 0   K obtains everywhere—at least, everywhere that g t t r , θ , ϕ > 0 or equivalently that g t t r , θ , ϕ 1 / 2 < . In the Schwarzschild (spherically-symmetrical non-rotating) special case, wherein g t t r = 1 r S r g t t r 1 / 2 = 1 r S r 1 / 2 , for a black hole this incorrect implication would pertain everywhere in the region r r S except at exactly  r = r S ; and for a non-black hole, whose radius exceeds r S , this incorrect implication would pertain everywhere without exception. Thus the correct implication is that any gravitator—black hole or non-black hole—must radiate: a black hole surrounded by a vacuum colder than T H , r and a non-black hole surrounded by a vacuum colder than T T , r cannot be in thermodynamic equilibrium with that vacuum, but must radiate into that vacuum and completely evaporate into that vacuum within a finite time! Thus at least the qualitative fact that Hawking (Tolman!) radiation emanates from all gravitators—not only from black holes but also from non-black holes—(even if not also quantitative values of T T , r and T H , r [16,43–49]) was discovered by Tolman [8,9], bolstered with Ehrenfest [10], at least as early as 1930! Any gravitator—black hole or non-black hole—surrounded by a 0   K vacuum cannot be at thermodynamic equilibrium unless it is enclosed within an opaque thermally insulating shell [64,65]   7 and thereby insulated from that vacuum: otherwise it will completely Hawking- (Tolman!-) evaporate into that vacuum within a finite time! This has been corroborated by recent research [12].
Black holes evaporate ever more rapidly and get hotter as they lose mass, hence completely evaporating into a vacuum at absolute zero ( 0   K ) within a finite time Δ t BH , evap . For evaporation of a Schwarzschild black hole into a 0   K vacuum [43–49]:
d M d t = 1 c 2 d E d t = 1 c 2 A σ T H , r 4 = 1 c 2 4 π r S 2 σ c 3 8 π G k M 4 = 1 c 2 4 π 2 G M c 2 2 σ c 3 8 π G k M 4 = σ 4 c 6 256 π 3 G 2 k 4 M 2 C 1 M 2 d t = M 2 C 1 d M Δ t BH , evap = d t = M initial 0 M 2 d M C 1 = 0 M initial M 2 d M C 1 = M initial 3 3 C 1 8.4116 × 10 17 M initial 1 kg 3 s 2.0972 × 10 67 M initial M 3 y ,
where the minus signs account for the black hole’s mass M decreasing during evaporation, A = 4 π r S 2 is its (decreasing) surface area, σ = 5.670374419 × 10 8 W / m 2   K 4 is the Stefan-Boltzmann constant [66],
C 1 σ 4 c 6 256 π 3 G 2 k 4 3.9628 × 10 15 kg 3 s 1 1.5905 × 10 68 M 3 y 1 M 3 6.2873 × 10 67 y ,
and M 1.9885 × 10 30 kg is the mass of the Sun [67]. The dot-equal sign (≐) means very nearly equal to.
By contrast, non-black holes evaporate ever more slowly and get cooler as they lose mass. But the time rate of this slowdown is itself sufficiently slow—they get cooler sufficiently slowly—that they, too, completely evaporate into a vacuum at absolute zero ( 0   K ) within a finite time Δ t NBH , evap . For a weak-field ( M initial r S c 2 / 2 G r initial r S = 2 G M / c 2 ) Schwarzschild (spherically-symmetrical non-rotating) non-black hole, g t t r and therefore also g t t r 1 / 2 is essentially constant at unity, and hence also by Equations (4)–(6) the Tolman [8–10] temperature T T r is essentially constant at T T , r , as M decreases from M initial to M final = 0 and r decreases from r initial to r final = 0 during the entire evaporation process. Moreover, A = 4 π r 2 r = A / π 1 / 2 / 2 , and, assuming uniform density ρ for simplicity (justified in the weak-field ( r NBH r S ) limit because gravity is too weak to significantly compress material with depth), also M = 4 π ρ r 3 / 3 = ρ r A / 3 = ρ A 3 / 2 / 6 π 1 / 2 A = π 1 / 3 6 M / ρ 2 / 3 . Hence in the weak-field ( r NBH r S ) limit for evaporation of a Schwarzschild (spherically-symmetrical non-rotating) uniform-density non-black hole into a 0   K vacuum:
d M d t = 1 c 2 d E d t = σ A T T , r 4 c 2 = π 1 / 3 σ T T , r 4 c 2 6 M ρ 2 / 3 C 2 M 2 / 3 d t = M 2 / 3 d M C 2 Δ t NBH , evap = d t = M initial 0 M 2 / 3 d M C 2 = M initial 0 M 2 / 3 d M C 2 = 3 M initial 1 / 3 C 2 ,
where the minus signs account for the non-black hole’s mass M decreasing during evaporation, and
C 2 π 1 / 3 6 / ρ 2 / 3 σ T T , r 4 c 2 3.0511 × 10 24 ρ 2 / 3 T T , r 4 kg 1 / 3 s 1 .
Δ t NBH , evap is finite because although d M / d t decreases with decreasing M, it does so only proportionately to M 2 / 3 . (In order to render Δ t NBH , evap infinite, d M / d t would have to decrease with decreasing M at least proportionately to M itself.) That Δ t NBH , evap is finite is corroborated by recent research [12].
Equations (11) and (12) yield an exact numerical value for C 1 . By contrast (even though the exact numerical values of all factors in C 2 except T T , r  are known) Equations (4)–(6), (13), and (14) do not yield enough information to provide an exact (or even less-than-exact) numerical value for C 2 and hence also for T T , r —albeit, as we showed in the eighth paragraph of this Section 3, they do yield enough information to prove that T T , r  must be finitely higher than absolute zero ( 0   K ). However, if our conjecture as per Equations (7)–(10) and the associated discussions is correct, then applying our result for T T , r in Equation (9) into Equation (13) does yield, at least for weak-field ( r NBH r S ) uniform-density Schwarzschild (spherically-symmetrical non-rotating) non-black holes, exact numerical values for C 2 , hence also for T T , r , and thence also for Δ t NBH , evap as per:
Δ t NBH , evap = 3 M initial 1 / 3 C 2 = 3 M initial 1 / 3 π 1 / 3 6 / ρ 2 / 3 σ T T , r 4 c 2 = 3 M initial 1 / 3 c 2 π 1 / 3 6 / ρ 2 / 3 σ T T , r 4 = 3 M initial 1 / 3 c 2 π 1 / 3 6 / ρ 2 / 3 σ T H , r r S r NBH 2 = 3 M initial 1 / 3 c 2 π 1 / 3 6 / ρ 2 / 3 σ c 3 8 π G k M r S r NBH 2 = 24 π 2 / 3 G k M initial 4 / 3 c σ ρ 6 2 / 3 r NBH r S 2 8.0141 ρ 2 / 3 M initial 4 / 3 r NBH r S 2 s .
Comparing Equations (11) and (13) with the same  M initial for both a Schwarzschild black hole and a weak-field ( r NBH r S ) uniform-density Schwarzschild non-black hole:
Δ t NBH , evap Δ t BH , evap = 3 M initial 1 / 3 C 2 M initial 3 3 C 1 = 9 C 1 C 2 M initial 8 / 3 C 3 M initial 8 / 3 .
Hence, if our conjecture as per Equations (7)–(10) and (15), and the associated discussions, is correct, then applying Equations (7)–(9), for weak-field ( r NBH r S ) uniform-density Schwarzschild non-black holes T T , r = T H , r r NBH / r S 2 , thereby yielding, at least for weak-field ( r NBH r S ) uniform-density Schwarzschild non-black holes, an exact numerical value for C 2 and thence also for C 3 :
C 2 3.0511 × 10 24 T H , r 4 ρ 2 / 3 r NBH r S 8 kg 1 / 3 s 1 6.9135 × 10 68 ρ 2 / 3 M initial 4 r NBH r S 8 kg 1 / 3 s 1
and
C 3 9 C 1 C 2 1.1689 × 10 40 ρ 2 / 3 T H , r 4 r S r NBH 8 kg 8 / 3 5.1587 × 10 53 ρ 2 / 3 M initial 4 r S r NBH 8 kg 8 / 3 .
Whether or not our conjecture is correct, with respect to Equation (16), this qualitative evaluation is valid: If M initial < C 3 3 / 8 , Δ t NBH , evap > Δ t BH , evap ; if M initial = C 3 3 / 8 , Δ t NBH , evap = Δ t BH , evap ; if M initial > C 3 3 / 8 , Δ t NBH , evap < Δ t BH , evap .
Thus all gravitators—black holes and non-black holes—are enveloped by atmospheres of equilibrium blackbody radiation. Because both T H , r > 0   K [16,43–49] and T T , r > 0   K [8–10], neither a black hole nor a non-black hole can be in thermodynamic equilibrium with a surrounding vacuum at (or at least sufficiently close to) absolute zero ( 0   K ), but must radiate into that vacuum and completely evaporate into that vacuum within a finite time, unless shielded from that vacuum by enclosure within an opaque thermally-insulating shell [64,65].   7
The Tolman-Hawking evaporation of a Schwarzschild black hole into a vacuum colder than T H , r and of a non-black hole into a vacuum colder than T T , r is in accordance with the Second Law of Thermodynamics. The entropy of a black hole is large [43–49], but the entropy of the radiation dispersed into a vacuum colder than T H , r by its Hawking-evaporation is even larger. The entropy of a non-black hole is not as large as that of a black hole of the same mass, affording even more scope for entropy to increase as it Tolman-evaporates into a vacuum colder than T T , r .
Tolman was aware of the concept of black holes (even if not of the moniker “black hole”): see, for example, the last paragraph of Section 96 of Ref. [9]. Yet nowhere does this enter into Tolman’s [8,9] derivations, bolstered with Ehrenfest [10], that at thermodynamic equilibrium temperature increases downwards   6 in any static, or even stationary, gravitational field [63]. Indeed, despite early contemplations of the concept of black holes [68–72], this concept [68–72] (and the moniker “black hole” [68–72]) was not mainstream until the 1960s [68–72]. Hence if Tolman’s [8,9] discovery, bolstered with Ehrenfest [10], had borne fruit circa 1930 (or shortly thereafter), it would have (i) initially been construed with respect to non-black holes and (ii) dubbed Tolman radiation rather than Hawking radiation: Hawking radiation would then initially have been construed as emanating from non-black holes—and dubbed Tolman radiation rather than Hawking radiation!
A caveat pertaining to Hawking-Tolman evaporation in general and to Tolman evaporation of Schwarzschild non-black holes in particular: Our results in this paper in general and in this Section 3 in particular relate, respectively, to gravitators in general and Schwarzschild non-black holes in particular that are bound solely by their own gravity. With respect to Schwarzschild non-black holes: any additional, non-gravitational, e.g., chemical, bonding, is neglected. Any additional, non-gravitational, e.g., chemical, bonding abetting gravitational bonding must necessarily increase Δ t NBH , evap , perhaps even rendering Δ t NBH , evap = if the temperature is absolute zero ( 0   K ). But this is a peripheral issue, not the central issue. The central issue relates only to gravitational bonding: to gravitators in general, and Schwarzschild non-black holes in particular, that are bound solely by their own gravity.
At this point, it is worthwhile to note the similarities   8 —owing to the equivalence principle [73,74]   8 —between Hawking (Tolman) radiation and Unruh radiation; but also a caveat.   8
These topics, and related ones, will be further discussed in Sections 4, 5, 6, and 7.

4. All Firewalls Are at the Planck Temperature

In Section 4, it may be helpful to envision a Schwarzschild black hole enclosed concentrically within an opaque thermally-insulating spherical shell at r shell [64,65].   7 Hawking radiation at temperature T H , r as per Equation (2) is reradiated and/or reflected downwards   6 from the inner surface of this spherical shell, suffering increasing gravitational blueshift with decreasing r in accordance with Equations (3)–(6) [8–11,16–18,43–49,64,65]. Since thermodynamic equilibrium obtains perfectly within the shell [64,65],   7 the caveat “(at least essentially)” can be deleted from the sentence containing Equation (2): radiation within the shell is exactly blackbody [64,65]. Indeed, enclosure of any radiation—whether emanating from a source or freely existing in space   9 —within an opaque thermally-insulating shell ensures perfect thermodynamic equilibrium and hence an exactly Planckian blackbody spectrum [64,65] (even if not immediately upon enclosure, then after the relaxation time). For example, if the Sun was so enclosed, the currently approximately blackbody radiation [67,70–78] at its photosphere would become exactly blackbody [64,65]. Without enclosure within an opaque thermally-insulating shell, radiation can be exactly blackbody; with enclosure, it must be exactly blackbody [64,65]. (Of course, exactly blackbody radiation incorporates the cutoff of the Planckian blackbody spectrum for wavelengths exceeding the size of an enclosure or cavity [79,80]. But this is not a consideration for our spherical shell, because it is at r shell [64,65].   7 ) Moreover, it should be noted that the Planckian form of any exactly-blackbody spectrum, and thus its having an exactly well-defined temperature, survives gravitational frequency shifting [81]—and also motional Doppler frequency shifting [81], cosmological frequency shifting [81], and any combination of any two or all three types of frequency shifting [81].
Prima facie, by Equation (3), it might seem that arbitrarily close to the Schwarzschild radius r S (but still at r > r S ) T H , r r S . But this is not so. Thus far, we have not taken into account that, if at r > r S , it is not possible, even in principle (let alone in practice) to be arbitrarily close to r S , because owing to quantum fluctuations spacetime breaks down as ruler distance [13] on the order of the Planck length [14–16]
l Planck = G c 3 1 / 2 = 1.616255 18 × 10 35 m
is approached. (The standard uncertainty is 0.000018 × 10 35 m [16].) Thus, even in principle (let alone in practice), it is not possible, if at r > r S , to be any closer to r S than at minimum radial ruler distance [14–16]
δ l min = l Planck = G c 3 1 / 2 = 1.616255 18 × 10 35 m
beyond r S .
We now derive T H , r as a function of radial ruler distance [13] δ l beyond r S , which we denote as T H , r S + δ l . We focus on regions just barely beyond r S , i.e., where δ l r S . Also, let δ r = r r S be Schwarzschild-coordinate radial distance [13] (which is also radial area distance and radial distance from apparent size [13]) beyond r S . We focus on regions just barely beyond r S , i.e., where, also, δ r r S . Obviously
1 r S r = 1 r S r S + δ r = r S + δ r r S r S + δ r = δ r r S + δ r = δ r r S δ r r S .
Applying Equations (1) and (21) [13],
d l = 1 r S r 1 / 2 d r d δ l = 1 r S r 1 / 2 d δ r = δ r r S δ r r S 1 / 2 d δ r = r S δ r 1 / 2 d δ r .
The last step of Equation (21) and the second-to-last step of Equation (22) are justified because we focus on regions just barely beyond r S , where δ r r S . Applying Equation (22), if δ r r S :
d δ l = δ r r S r S δ r 1 / 2 d δ r δ l = δ r r S 0 δ r r S δ r 1 / 2 d δ r = 2 r S δ r 1 / 2 δ r δ l r S δ r = δ r δ l r S δ l 2 4 r S r S δ r 1 / 2 = δ r δ l r S δ l 2 δ r = δ l 2 δ l 2 4 r S = 2 r S δ l .
Hence, applying Equations (2)–(6), (21), (22), and (23), if δ r δ l r S :
T H , r = T H , r S + δ r = δ r δ l r S T H , r r S δ r 1 / 2 T H , r S + δ l = δ r δ l r S 2 T H , r r S δ l = 2 c 4 π k r S r S δ l = c 2 π k δ l .
As noted in the paragraph containing Equations (19) and (20), even in principle (let alone in practice), δ l can be no smaller than δ l min = l Planck [14–16]. Thus, minimizing δ l at δ l min = l Planck , by Equations (19), (20), and (24) we obtain
T firewall = T H , r S + δ l min = T H , r S + l Planck = c 2 π k δ l min = c 2 π k l Planck = c 2 π k G c 3 1 / 2 = 1 2 π k c 5 G 1 / 2 = 1 2 π 1 k c 5 G 1 / 2 = T Planck 2 π ,
where
T Planck = 1 k c 5 G 1 / 2 = 1.416784 16 × 10 32   K
is the Planck temperature [19]. (The standard uncertainty is 0.000016 × 10 32   K [19].)
This result is independent of the mass M and hence also of the Schwarzschild radius r S = 2 G M / c 2 of a Schwarzschild black hole. As M and hence also r S increases, by Equation (2) T H , r decreases in inverse proportion. But r S / δ l for any given δ l r S in general and hence r S / δ l min = r S / l Planck in particular increases in direct proportion. Hence in accordance with Equations (2) and (24)–(26) these two opposing factors cancel out. Because of quantum fluctuations in the metric at length scales of l Planck [14–16], Equation (25) may be pushing the limit of accuracy of Equation (24), but we should expect Equation (25) to be valid at least in some average sense. Accordingly, perhaps we should not be too adamant about the small numerical factor of 1 / 2 π in Equation (25), and hence recapitulate Equation (25) as
T firewall = T H , r S + l Planck T Planck .
By Equation (24), recapitulated with the help of
c 2 π k = 0.0003644464403 m   K
as
T H , r S + δ l = δ r δ l r S c 2 π k δ l = 0.0003644464403 m   K δ l ,
T H , r S + δ l still has high values in the region l Planck δ l r S , hence with quantum fluctuations in the metric of Equation (1) being negligible [14–16]. For example, the temperature of the Sun’s core, 1.571 × 10 7   K [67], is equaled at δ l = 2.3198 × 10 11 m , i.e., only slightly less than typical atomic dimensions; the (effective [67,75–78]) temperature of the Sun’s photosphere, 5772   K [67], is equaled at δ l = 6.3140 × 10 8 m , i.e., the dimensions of small microbes; and room temperature, 300   K , is equaled at δ l = 1.2148 × 10 6 m , less than two orders of magnitude below the limit of naked-eye visibility 10 4 m .
Now let us consider the ruler-distance [13] wavelength of Hawking radiation in the region only slightly beyond r S , i.e., where δ r δ l r S . The ruler-distance [13] wavelength λ r S + δ l Wien , max of blackbody radiation in general and of Hawking radiation in particular at the Wien’s-Displacement-Law maximum with respect to wavelength [66,82] corresponding to temperature T is [66,82]
λ Wien , max = 0.002897771955 T m .
(Since we are focusing on wavelength, we employ the Wien’s-Displacement-Law maximum with respect to wavelength [66,82] as opposed to that with respect to frequency [66,82].) Hence by Equations (28) and (30) [66,82]:
λ Wien , max = 0.002897771955 T m c 2 π k c 2 π k = 0.002897771955 T m 0.0003644464403 m   K c 2 π k = 0.002897771955 m   K 0.0003644464403 m   K c 2 π k T = 1.265466416 c k T λ r S + δ l Wien , max = δ r δ l r S 1.265466416 c k T H , r S + δ l = 1.265466416 c k c 2 π k δ l = 1.265466416 × 2 π δ l = 7.951159991 δ l .
The numerical factor 1.265466416 × 2 π = 7.951159991 is dimensionless and hence is valid in any self-consistent system of units. In the third line of Equation (31) we applied the second line of Equation (24). In accordance with the reasoning concerning quantum fluctuations in the metric in the paragraph ending with Equation (27) [14–16], perhaps we should not be too adamant about the small numerical factor of 1.265466416 × 2 π = 7.951159991 in the last term of Equation (31), and hence recapitulate Equation (31) as
λ r S + δ l Wien , max δ r δ l r S δ l .
Thus the ruler-distance [13] wavelength λ r S + δ l Wien , max of Hawking radiation in the region δ r δ l r S is on the order of the ruler distance [13] δ l itself. In particular, at δ l min = l Planck [14–16]
λ r S + δ l min Wien , max = λ r S + l Planck Wien , max l Planck .
Hawking-radiation photons for which Equations (25)–(27) and (33) apply, and consequently for which T firewall T Planck and thus E = h ν k T firewall k T Planck E Planck = m Planck c 2 [14–16,83–86], are themselves Planck-mass black holes [14–16,87–89], specifically, Planck-mass geons [87–89], and thereby themselves contribute to the breakdown of spacetime as the Planck scale is approached, i.e., as δ l δ l min = l Planck [14–16,87–89].
In accordance with the three immediately preceding paragraphs, and for consistency with Equation (31) keeping the numerical factor 7.951159991 [66,82], by Equations (30) and (31) [66,82]
λ Wien , max = 0.002897771955 T m = δ r δ l r S 7.951159991 δ l .
Thus Hawking radiation with λ Wien , max corresponding to values of T that are still high occurs in the region l Planck δ l r S , hence with quantum fluctuations in the metric of Equation (1) being negligible [14–16]. For example, λ Wien , max corresponding to the temperature of the Sun’s core, 1.571 × 10 7   K [67], is equaled at δ l = 2.3198 × 10 11 m , i.e., only slightly less than typical atomic dimensions; λ Wien , max corresponding to the (effective [67,75–78]) temperature of the Sun’s photosphere, 5772   K [67], is equaled at δ l = 6.3140 × 10 8 m , i.e., the dimensions of small microbes; and λ Wien , max corresponding to room temperature, 300   K , is equaled at δ l = 1.2148 × 10 6 m , less than two orders of magnitude below the limit of naked-eye visibility 10 4 m .
Of course, the last two lines of Equation (31), and Equations (32) and (34) [let alone Equation (33)], do not apply in the region r r S . For, as r , δ l r r S + r S / 2 ln r / r S [90], whilst applying Equation (2) and the first two lines of Equation (31) [66,82]:
λ r Wien , max = 1.265466416 c k T H , r = 1.265466416 c k c 4 π k r S = 1.265466416 × 4 π r S = 15.90231998 r S = constant .
The numerical factor 1.265466416 × 4 π = 15.90231998 is dimensionless and hence is valid in any self-consistent system of units.
We have considered Schwarzschild black holes whose only energy source is their own Hawking radiation. This may eventually be the case for actual black holes if the Universe expands forever. But in the current Universe, black holes are bathed by photons emanating from r r S —effectively from r —far more energetic than thermal photons at temperature T H , r as per Equation (2): photons from the T CBR = 2.725   K [50] cosmic background radiation [50], from starlight, etc. [50]. Radiation comprised of these far more energetic photons will be blueshifted to T firewall T Planck as given by Equations (25)–(27) at r S + δ l with δ l l Planck . But photons corresponding to T firewall T Planck , i.e., for which E = h ν k T firewall k T Planck E Planck = m Planck c 2 [14–16,87–89], are themselves Planck-mass black holes [14–16,87–89], specifically, Planck-mass geons [87–89], and thereby themselves might contribute to the breakdown of spacetime at this  r S + δ l , i.e., at this  δ l l Planck , well before  δ l min = l Planck is approached [14–16,87–89]. Hence in the current Universe we should consider at least the possibility of the breakdown of spacetime at this  r S + δ l , i.e., at this  δ l l Planck , well before  δ l min = l Planck is approached [14–16,87–89]. But this is not what we mean by a Schwarzschild black hole’s firewall. By a Schwarzschild black hole’s firewall we mean that which is intrinsic to the black hole itself, i.e., owing solely to its own Hawking radiation.

5. The Exponential Nature of the Gravitational Frequency Shift

In Section 5, as in Section 4, it may be helpful to envision a Schwarzschild black hole enclosed concentrically within an opaque thermally-insulating spherical shell at r shell [64,65].   7 Hawking radiation at temperature T H , r as per Equation (2) is reradiated and/or reflected downwards   6 from the inner surface of this spherical shell, suffering increasing gravitational blueshift with decreasing r in accordance with Equations (3)–(6) [8–11,16–18,43–49,64,65].
Expressed in terms of r, at r r S the relativistic gravitational scalar potential Φ of a Schwarzschild black hole and its magnitude Φ are [13,17,18]
Φ = c 2 2 ln 1 2 G M r c 2 = c 2 2 ln 1 r S r Φ = c 2 2 ln 1 2 G M r c 2 = c 2 2 ln 1 r S r .
Applying Equations (2), (3), (21), (22), and (23) [especially Equation (21) and the last two lines of Equation (23)], if δ r δ l r S , expressing Φ and Φ in terms of δ r and δ l [13,17,18]:
Φ = δ r δ l r S c 2 2 ln δ r r S = c 2 2 ln δ l 2 4 r S 2 = c 2 2 ln δ l 2 r S 2 = c 2 ln δ l 2 r S Φ = δ r δ l r S c 2 2 ln r S δ r = c 2 2 ln 4 r S 2 δ l 2 = c 2 2 ln 2 r S δ l 2 = c 2 ln 2 r S δ l .
It may be interesting to note that corresponding to minimum-definable ruler distance [13] δ l min = l Planck [14–16] beyond r S
Φ r S + δ l min = Φ r S + l Planck = c 2 ln 2 r S δ l min = c 2 ln 2 r S l Planck = c 2 ln 2 r S G c 3 1 / 2 = c 2 ln r S 4 c 3 G 1 / 2 = c 2 ln 4 r S 2 c 3 G 1 / 2 = c 2 2 ln 4 r S 2 c 3 G = c 2 2 ln A c 3 π G = c 2 2 ln 16 G M 2 c = c 2 2 ln 4 S π k ,
where A is the surface area of a black hole and S is its entropy [43–49].
We re-emphasize that a relativistic gravitational scalar potential Φ and hence also its magnitude Φ [17,18], and the relation thereof to gravitational potential energy [17,18], are valid concepts for all static, and even stationary, spacetimes [63] (not just Schwarzschild spacetime [91–93]). And that the spacetime engendered at r r S by any Schwarzschild black hole and at all  r 0 by any Schwarzschild non-black hole (we focus on Schwarzschild metrics) is static [91–93], not merely stationary [63].
The blueshift of any photon (Hawking/Tolman-radiation photon or otherwise) whose frequency, energy, and mass [83–86] at r are ν r , E   r = h ν r , and m   r = E   r / c 2 = h ν r / c 2 [83–86], respectively, upon falling radially inwards from r , increases exponentially rather than merely linearly with decreasing Φ (or, equivalently, with increasing Φ ) [17,18], in accordance with [17,18]
ν Φ = ν r e Φ / c 2 E Φ = h ν Φ = E   r e Φ / c 2 = h ν r e Φ / c 2 m Φ = E Φ c 2 = h ν Φ c 2 = m   r e Φ / c 2 = E   r e Φ / c 2 c 2 = h ν r e Φ / c 2 c 2 .
This obtains because as a photon falls and gets blueshifted its mass [83–86] m = E / c 2 = h ν / c 2 [which of course is solely its (kinetic energy) / c 2 [83–86], because a photon’s rest mass is zero [83–86]] increases: the photon gets more massive as it falls. Thus as a photon falls through successive ruler-distance [13] increments d l , a Schwarzschild black hole’s gravitational field at r r S , a Schwarzschild non-black hole’s gravitational field at all  r 0 —indeed, the gravitational field G = d Φ / d l in any static, or even stationary, spacetime [63,91–93]—does successive increments of (positive) work [92,93]
d W = m d Φ = m d Φ d l d l = m G d l
not on a fixed mass m but on an ever-increasing mass m. [The minus sign in G = d Φ / d l obtains because G acts downwards,   6 i.e., in the direction of decreasing l. d W in Equation (40) is positive, because G is negative, and both d Φ and d l are negative during infall.] d W / d Φ and thus the rate of increase of m = E / c 2 = h ν / c 2 with decreasing Φ is proportional to m = E / c 2 = h ν / c 2 itself: consequently the exponential form of Equation (39).
Hence also, in accordance with Equations (3), (24), (25), (36), (37), (39), and (40), the temperature T of any Planckian blackbody distribution of photons increases exponentially rather than merely linearly with decreasing Φ (or, equivalently, with increasing Φ ) [16–18,43–49,64,65,81,83–86].
Of course, the same reasoning also applies in reverse: as a photon rises a Schwarzschild black hole’s gravitational field at r > r S —indeed, the gravitational field G = d Φ / d l in any static, or even stationary, spacetime [63,91–93]—does negative work on, or equivalently receives positive work from, not a fixed mass m but an ever-decreasing mass m. d W / d Φ and thus the rate of decrease of m = E / c 2 = h ν / c 2 is proportional to m = E / c 2 = h ν / c 2 itself: consequently as per Equations (39) and (40) a rising photon’s mass [83–86] m = E / c 2 = h ν / c 2 decreases exponentially rather than merely linearly with increasing Φ (or, equivalently, with decreasing Φ ) [16–18,43–49,64,65,81,83–86]. Hence also, in accordance with Equations (3), (24), (25), (36), (37), (39), and (40), the temperature T of any Planckian blackbody distribution of photons decreases exponentially rather than merely linearly with increasing Φ (or, equivalently, with decreasing Φ ) [16–18,43–49,64,65,81,83–86].
By contrast, for a slowly radially-moving (slow physicalnot necessarily slow coordinate—radial velocity V phys = d l / d τ c ) nonzero-rest-mass particle, the increase of total mass in free fall (and its decrease in free rise from an upwards   6 flying start) is on a pro rata basis much smaller than for a photon—a linear rather than exponential function of Φ (or Φ ). This obtains because its (kinetic energy) / c 2 = V 2 / 2 c 2 is only a negligibly small fraction of its total mass—not the entirety [83–86] of its total mass as is the case for a photon (or other zero-rest-mass particle) [83–86].
Of course, the First Law of Thermodynamics (energy conservation) always obtains. The kinetic energy that any entity gains (loses) by falling (rising) in a gravitational field is exactly offset by the energy of the gravitational field itself becoming more (less) strongly negative. This point will be discussed more thoroughly in Section 6.
In wrapping up Section 5, we note that for static, and even stationary, spacetimes [63], the relativistic gravitational scalar potential Φ is related to the time-time component of the metric in accordance with [94]
g t t = e 2 Φ / c 2 Φ = c 2 2 ln g t t .
Hence in static, and even stationary, spacetimes [63], the substitution Φ c 2 2 ln g t t can be made in Equations (36)–(40).

6. Negative Gravitational Mass-Energy and Birkhoff’s Theorem versus Massiveness of Firewalls

Thus far, we have taken for granted that a firewall does not contribute (at a maximum, not more than negligibly) to the mass M of a Schwarzschild black hole. But this has been seriously questioned [3]. It has been averred that this cannot be even approximately true for any Schwarzschild black hole whose mass M appreciably exceeds the Planck mass m Planck = c / G 1 / 2 [3]—a minimum-possible-mass M = M min = m Planck = c / G 1 / 2 Schwarzschild black hole—which would Hawking-evaporate on a time scale on the order of the Planck time t Planck = G / c 5 1 / 2 [3]. And that at best this can just barely be even approximately true even if  M = M min = m Planck = c / G 1 / 2 [3]. This is the firewall-mass problem [3].
There is not universal agreement concerning the firewall-mass problem [3]. Counter-arguments resolving this problem have been proposed [4].
In Section 6, we do not make any assumption about what the mass of a firewall might be: small, large, or perhaps annulled to zero (except that it is not negative) [3,4]. However, we consider the firewall-mass problem [3], and propose an at least prima facie tentative resolution thereto. Our tentative resolution is based on: (i) the mass of a firewall (whatever it might be, if not otherwise annulled to zero [4]) being exactly canceled by the negative gravitational mass [17,18] = ( negative gravitational energy ) / c 2 [17,18] accompanying its formation, (ii) the unchanged observations of a distant observer upon formation of a firewall, and (iii) Birkhoff’s Theorem [20–31]—actually first discovered by Jørg Tofte Jebsen [25–28]. This is in addition to, and perhaps may complement, other lines of reasoning [4] disputing massiveness [3] of firewalls. (There is a caveat [28–31]   1 with respect to Birkhoff’s Theorem [20–31], but it [28–31]   1 is not relevant with respect to our considerations.   1 )
The viewpoint [3] that formation of a firewall imparts a huge net increase to the mass of a Schwarzschild black hole [3] seems to overlook the negative gravitational mass [17,18] = ( negative gravitational energy ) / c 2 [17,18] contribution to the black-hole/firewall system. The negativity of gravitational energy [17,18] is the perhaps the central aspect of our tentative resolution of the firewall-mass problem [3]. We hope to show that the negative gravitational energy [17,18] accompanying formation of a firewall exactly—not merely approximately—cancels the firewall mass, so that the mass M of a black hole remains exactly—not merely approximately—unchanged if a firewall forms. Is this, at least prima facie, what Ref. [3] overlooks? Reference [3] derives the mass of an already-extant firewall of an already-collapsed black hole, but seems to overlook the increased negativity of gravitational mass-energy accompanying formation of the firewall during collapse.
We note that the negative gravitational mass-energy accompanying formation of a firewall should not be confused with considerations regarding negative energy states of the firewall itself [3]. We do not make any assumption about what the mass of a firewall might be—except that in accordance with the first two paragraphs of the section entitled “Discussion” in Ref. [3], we always construe its mass (if not annulled to zero [4]) to be positive—even if there exist negative energy states: the squares of both positive and negative numbers are positive: see the term E F 2 in Equation (10) of Ref. [3]. We show that, whatever the mass of a firewall might be, the negative gravitational mass [17,18] = ( negative gravitational energy ) / c 2 [17,18] accompanying its formation annuls it (even if it is not otherwise annulled [4])—effecting zero net change in the mass of a Schwarzschild black hole.
Consider a spherically-symmetrical non-rotating gravitator of mass M but of sufficiently low average density that it is a Schwarzschild non-black hole ( r > r S = 2 G M / c 2 M < r S c 2 / 2 G ), surrounded by a vacuum at absolute zero ( 0   K ). As shown in Section 3, this gravitator will completely Tolman-radiation [8–10] evaporate within a finite time (see also Garrod [11]), yielding energy E = M c 2 to a distant observer at r obs r > r S . We re-emphasize that this has been corroborated by recent research [12].
Now instead consider another identical spherically-symmetrical non-rotating gravitator of mass M. But this time let the structural strength of the gravitator be annulled, so that it gravitationally collapses radially to a Schwarzschild black hole. This gravitator will then completely Hawking-radiation evaporate within a finite time, also yielding the same energy E = M c 2 to a distant observer at r obs r r S . Indeed, this is required not only by the First Law of Thermodynamics (energy conservation), but also by Birkhoff’s Theorem [20–31]. (There is a caveat [28–31]   1 with respect to Birkhoff’s Theorem [20–31], but it [28–31]   1 is not relevant with respect to our considerations.   1 ) For Birkhoff’s Theorem [20–31] states that any purely radial gravitational collapse (or any purely radial dispersion against gravity from a flying start) of a spherically-symmetrical non-rotating gravitator cannot cause any change detectable by a distant observer [not even gravitational waves, because radial collapse (or radial dispersion) does not generate them [20–31]]: Birkhoff’s Theorem [20–31] authorizes no exception for gravitational collapse of the innermost shell of a gravitator’s Tolman-Hawking [8–11] radiation atmosphere to a firewall. This is possible if and only if the mass of the gravitator does not change during collapse—even if a firewall forms. And this, in turn, is possible if and only if the mass of the firewall is exactly counterbalanced by the increased negativity of gravitational mass-energy accompanying its formation.
Thus there must be zero net change in mass of the gravitator. Any increase in mass—whether due to formation of a firewall and/or otherwise—accompanying collapse must be exactly counterbalanced by a negative contribution. Gravitational mass = (gravitational energy) / c 2  is always a negative contribution to mass. And the only possible counterbalancing negative contribution is the gravitational mass-energy of the gravitator becoming more strongly negative during collapse. This must be true whether or not a firewall forms. If a firewall does not form, the increase in mass of the collapsing gravitator’s Tolman-Hawking [8–11] radiation atmosphere will be less than if one does form—but so will the increase in the negativity of gravitational mass-energy.
It may be helpful to briefly expound on Tolman-Hawking [8–11] radiation atmospheres. Consider a spherically-symmetrical non-rotating entity (Schwarzschild black hole or Schwarzschild non-black hole) enclosed concentrically within an opaque thermally-insulating spherical shell at r shell [64,65].   7 Such an entity is enveloped by a Tolman-Hawking [8–11] radiation atmosphere. Because the entity is enclosed within an opaque thermally-insulating spherical shell, its Tolman-Hawking [8–11] radiation atmosphere is at thermodynamic equilibrium throughout. Hence photons of radiation emanate from anywhere in this radiation atmosphere. To be specific, if this entity is a black hole, it is equally valid to construe photons emanating either (i) from r S + l Planck and then suffering gravitational redshift upon streaming outwards towards the inner surface of our spherical shell at r or (ii) from the inner surface of our spherical shell at r and then suffering gravitational blueshift upon falling inwards. Thus either we can construe Hawking radiation as suffering maximal gravitational redshift at r and no gravitational redshift at r S + l Planck , or we can construe it as suffering no gravitational blueshift at r and maximal gravitational blueshift at r S + l Planck . This viewpoint is valid because: (a) the entire region within our spherical shell is at thermodynamic equilibrium throughout. And at thermodynamic equilibrium, the principles of microscopic reversibility and detailed balance obtain [95]: hence it is equally valid to consider microscopic processes occurring in either the “forward” or “reverse” direction [95]. Indeed, at thermodynamic equilibrium, which direction (i) or (ii) immediately above is construed as “forward” or “reverse” is arbitrary [95]. (There are caveats [96,97],   10 but they are not relevant with respect to our considerations.   10 ) (b) Curved spacetime is hot [8–11]. Thus—if the gravitational frequency shift and hence temperature increasing downwards   6 in gravitational fields is taken into account [8–11,81]—it is equally valid to construe Tolman-Hawking [8–11] radiation as emanating from any  r > r S [8–11,81,95]. A Tolman-Hawking [8–11] radiation photon of mass m   r = E   r / c 2 = h ν r / c 2 k T H , r / c 2 [83–86] at the inner surface of our spherical shell at r does indeed gain mass m Planck m   r during its infall to r S + l Planck , i.e., to δ l min l Planck [14–16,83–86], attaining mass m Planck = E Planck / c 2 = h ν r S + l Planck / c 2 k T firewall / c 2 k T Planck / c 2 after having fallen to r S + l Planck , i.e., to δ l min l Planck [14–16,83–86]. But the increase m Planck m   r in the photon’s mass [14–16,83–86] that occurs during its infall is exactly counterbalanced by the increased negativity of the gravitational mass-energy [17,18] of the black-hole/photon system [83–86] that, by the First Law of Thermodynamics (energy conservation), also occurs during the photon’s infall. Thus the net contribution to the mass of the black-hole/photon system [83–86] continues to be only  m   r —it does not increase by m Planck m   r to m Planck exactly as if the photon had not suffered infall!
If this is true with respect to any one infalling photon, then it must also be true with respect to all of the infalling photons combined required to produce a spherical shell of equilibrium blackbody radiation with inner boundary at r S , of ruler-distance [13] radial thickness l Planck , and at temperature T Planck —i.e., to produce a firewall. Hence at least prima facie it seems that a large increase in the mass [3]—indeed any increase in mass at all—of the black hole attributable to firewall formation [3] is exactly canceled out to zero.
We re-emphasize that the downwards   6 increase in the temperature of Tolman-Hawking [8–11] radiation in the gravitational fields of Schwarzschild gravitators (black holes and non-black holes) is a special case of the general result of relativistic thermodynamics that at thermodynamic equilibrium temperature increases downwards   6 in any gravitational field [8–11] (at least, in any static, or even stationary, one [63,91–93]). Tolman-Hawking [8–11] radiation should be construed as emanating not only from r S + l Planck —indeed not only from any  r r S + l Planck —in the gravitational field of a Schwarzschild black hole—but from anywhere in any gravitational field whatsoever. This was very well conveyed by a seminar given by Dr. James H. Cooke at the Department of Physics at the University of North Texas in the 1980s—and most succinctly expressed by the title of this seminar: “Curved spacetime is hot”—confirming Tolman [8–10] (see also Garrod [11], and recall our Sections 3, 4, and 5). Of course, by “hot” it is meant hotter than absolute zero ( 0   K )—in even the weakest gravitational fields. Tolman-Hawking [8–11] radiation emanates from every location in any gravitational field however weak in general—not only from black holes, but also from non-black holes: Curved spacetime is hot [at least, hotter than absolute zero ( 0   K )] in general. This is required for consistency with temperature increasing downwards   6 given thermodynamic equilibrium in any gravitational field, however weak [8–11]. In this regard we re-emphasize, as Dr. James H. Cooke pointed out, that not only black holes, but also non-black holes, Tolman-Hawking [8–11]) radiate: In this regard, it may at this point be worthwhile to again recall Section 3. Indeed, as we noted in Section 3, when Tolman [8,9], bolstered with Ehrenfest [10], anticipated Hawking radiation (see also Garrod [11]), if that anticipation had borne fruit circa 1930 (or shortly thereafter), it would have (i) initially been construed with respect to non-black holes and (ii) dubbed Tolman radiation rather than Hawking radiation!
Generalizing, the free fall of any entity in any gravitational field cannot result in any change in the mass of the gravitator/entity system, because by the First Law of Thermodynamics (energy conservation) the gain in the falling entity’s kinetic energy [via increased frequency if it is a photon, or via increased physical downwards   6 velocity V phys = d l / d τ (not necessarily increased coordinate downwards   6 velocity) if it is of nonzero rest mass] must be exactly counterbalanced by the gravitational mass-energy [17,18] of the gravitator/entity system becoming more strongly negative. And likewise the free rise (from an upwards   6 flying start) of any entity in any gravitational field cannot result in any change in the mass of the gravitator/entity system, because by the First Law of Thermodynamics (energy conservation) the loss in the rising entity’s kinetic energy [via decreased frequency if it is a photon, or via decreased physical upwards   6 velocity V phys = d l / d τ (not necessarily decreased coordinate upwards   6 velocity) if it is of nonzero rest mass] must be exactly counterbalanced by the gravitational mass-energy [17,18] of the gravitator/entity system becoming less strongly negative. Furthermore this remains true even if the fall or rise is not free but retarded by friction [21], because friction merely thermalizes the entity’s kinetic energy within the gravitator/entity system. For example, a landslide on Earth (whether or not retarded by friction) does not change Earth’s total mass-energy E Earth = M Earth c 2 (which includes the negative contribution from Earth’s gravitational energy), because the kinetic energy of the landslide (whether or not thermalized by friction [21]) is exactly counterbalanced by the gravitational mass-energy [17,18] of the Earth/landslide system becoming more strongly negative.
We briefly remark that Earth’s negative gravitational mass-energy [17,18] reduces Earth’s mass by a fraction on the order of V escape 2 / c 2 10 9 , where V escape is the escape velocity from Earth’s surface ( 1.1 × 10 4 m / s ). While this fraction is small in relative terms, in absolute terms it is a substantial negative contribution to Earth’s mass, on the order of the mass of an asteroid 10 km in diameter( 10 3 of Earth’s diameter)—e.g., the K–T boundary asteroid [98] that was the major factor (even if not the only one) that ended the dinosaurs’ reign [98].   11
We close Section 6 with this speculative paragraph. It has been speculated [3] that owing to a firewall perhaps an infalling particle “burns up at the horizon [3]”. So we are steered in the direction of asking the following four admittedly speculative questions: (i) Might the particle be saved from falling through the horizon, i.e., through the Schwarzschild radius r S of a black hole, by burning up? (ii) If so, does this at least prima facie seem to suggest the possibility that a collapsing near-black hole might be saved from falling through its own Schwarzschild radius r S by beginning to burn up mass as soon as its surface approaches a ruler distance [13] of one Planck length beyond r S ( r S + l Planck ), i.e., that black holes can thus come within a gnat’s eyelash of forming, but cannot completely form? This gnat’s eyelash would of course not be sufficient to result in any measurable or observable astronomical or astrophysical dissimilarity from completely-formed black holes. (iii) And, for example, given (ii) immediately above, that as Hawking evaporation of a gnat’s-eyelash near-black hole proceeds into a vacuum whose temperature is at (or at least sufficiently close to) absolute zero ( 0   K ), its surface always remains a ruler distance [13] of one Planck length beyond r S , this being maintained until Hawking evaporation is complete? (iv) Might this be relevant, for example, with respect to solving the black-hole information paradox? For, if black holes can come within a gnat’s eyelash of fully forming but cannot fully form, no information can ever fall into a fully-formed black hole and hence there is no need for it to be retrieved from one. Of course, various (hopefully, at least to some extent, mutually compatible) resolutions of the black-hole information paradox have been proposed [99–108]   12 . We note that if black holes can thus come within a gnat’s eyelash—but no further—of forming, the maximum possible depth Φ max of their gravitational wells is finite. For then, Φ max = Φ r S + δ l min = Φ r S + l Planck as per Equations (19) and (38).

7. Relativistic Gravitational Temperature Gradients Cannot Defy the Second Law of Thermodynamics

We note that equilibrium vertical gravitational temperature gradients that exist [8–11]—indeed that are required [8–11]—by relativistic thermodynamics [8–11] cannot be exploited to violate the Second Law of Thermodynamics.
First, consider a gravitator enclosed concentrically within an opaque thermally-insulating spherical shell.   7 Now consider a heat engine trying to exploit the equilibrium relativistic gravitational temperature gradient, via a hot reservoir at a lower altitude at temperature T hot and a cold reservoir at a higher altitude at temperature T cold .
Macroscopic consideration: Thermodynamic equilibrium [8–11,109,110] exists within the shell, and thermodynamic equilibrium [8–11,109,110] necessarily implies hydrostatic equilibrium [110–115] (but not necessarily vice versa [8–11,109–115]). Owing to hydrostatic equilibrium [110–115] that thermodynamic equilibrium [8–11,109,110] necessarily implies, the weight E g / c 2 of a parcel of thermal energy E where the gravitational acceleration is g [8–11] exactly counterbalances its tendency to flow from higher temperatures at lower altitudes to lower temperatures at higher altitudes, so macroscopically there is staticity and hence no flow of heat that a heat engine can utilize. [Likewise at hydrostatic equilibrium—even without, let alone with, thermodynamic equilibrium, and either relativistically or non-relativistically—the weight m g of a parcel of fluid (gas or liquid) of mass m where the gravitational acceleration is g [8–11] exactly counterbalances its tendency to flow from higher pressures at lower altitudes to lower pressures at higher altitudes, so macroscopically there is staticity and hence no flow of fluid that a pneumatic engine can utilize.]
Microscopic consideration: While macroscopically thermodynamic equilibrium is, or at least can be construed as, static, by contrast, microscopically, thermodynamic equilibrium is dynamic. At thermodynamic equilibrium, individual blackbody-radiation photons move up and down in any gravitational field. But: Even without our engine trying to convert any heat whatsoever from the hot reservoir into work, the gravitational redshift diminishes the temperature of equilibrium blackbody photons radiated at T hot from the lower altitude of the hot reservoir to T cold upon them reaching the higher altitude of the cold reservoir—thus diminishing the Carnot efficiency ϵ Carnot = 1 T cold / T hot to ϵ Carnot = 1 T cold / T cold = 0 . What the gravitational temperature gradient giveth, the gravitational redshift taketh away [8–11]: after the gravitational redshift has taken its cut, there is nothing left over to be converted into work [8–11]. (Similarly, in accordance with either relativistic or non-relativistic hydrodynamics and thermodynamics [109–115], even though microscopically at thermodynamic equilibrium individual fluid molecules comprising a gas or liquid move up and down in any gravitational field, gravitational pressure gradients cannot be exploited by a pneumatic engine: at hydrostatic equilibrium—even without, let alone with, thermodynamic equilibrium [109–115], and either relativistically or non-relativistically—what the gravitational pressure gradient giveth, the weight taketh away [110–115].)
Next, consider a gravitator not enclosed concentrically within an opaque thermally-insulating spherical shell, but instead surrounded by a vacuum at (or at least sufficiently close to) absolute zero ( 0   K ). Such a gravitator is not at thermodynamic equilibrium. Any gravitator’s Tolman-Hawking equilibrium blackbody radiation will disperse into the surrounding vacuum. Hence without enclosure within such a shell, a heat engine can operate—but only at the expense of the increase in entropy owing to dispersal of the radiation into the vacuum. [Similarly, without enclosure within a shell, a gravitator’s gas (liquid) atmosphere (hydrosphere) will evaporate into a surrounding vacuum. Hence without enclosure within a shell, a pneumatic engine also can operate—but only at the expense of the increase in entropy owing to dispersal of the atmosphere (hydrosphere) into the vacuum.]
Hence both with and without enclosure by an opaque thermally-insulating spherical shell, the Second Law of Thermodynamics is obeyed.
To re-emphasize, thermodynamic equilibrium [8–11,109,110] necessarily implies hydrostatic equilibrium [110–115], but not necessarily vice versa [8–11,109–115]. The terms “hydrostatic equation [110–113]” or “barometric equation [115]” are sometimes employed to denote hydrostatic equilibrium [110–113] but not necessarily thermodynamic equilibrium [8–11,109–115]. Earth’s atmosphere and oceans are typically at hydrostatic equilibrium (or at least very nearly so). But, of course, because they are impelled by the large temperature difference between the hot solar disk and the cold rest of the sky, they are not at thermodynamic equilibrium.

8. Conclusion

Following introductory remarks in Section 1, in Section 2 we reviewed basic concepts pertaining to Schwarzschild black holes and Hawking radiation. In Section 3 we discussed the anticipation of Hawking radiation—albeit from non-black holes—initially by R. C. Tolman and shortly thereafter with P. Ehrenfest (see also Garrod [11]). This has been corroborated by recent research [12]. This anticipation culminates in the proof [8–12] that any gravitator—black hole or non-black hole—must radiate and hence cannot be in thermodynamic equilibrium with a surrounding vacuum at (or at least sufficiently close to) absolute zero ( 0   K ), but must completely evaporate into that vacuum within a finite time. The times required for evaporation of black holes and non-black holes were compared. In Section 4, we showed that (i) if firewalls exist, they can originate via Hawking radiation at the minimum possible ruler distance [13] (the Planck length [14–16]) beyond the Schwarzschild horizon, where it has not suffered any gravitational redshift [17,18], or, alternatively, suffered maximal gravitational blueshift and (ii) the firewall temperature is on the order of the Planck temperature [19], independently of the mass and hence also of the Schwarzschild radius of a Schwarzschild black hole. In Section 5, we explained the exponential nature of the gravitational frequency shift as a function of the gravitational potential. In Section 6, we considered the firewall-mass problem [3], and provided an at least prima facie tentative resolution thereto based on: (i) the mass of a firewall being canceled by the negative gravitational mass [17,18] = ( negative gravitational energy ) / c 2 [17,18] accompanying its formation, (ii) the unchanged observations of a distant observer upon formation of a firewall, and (iii) Birkhoff’s Theorem [20–31]. (The basis upon Birkhoff’s Theorem is unaltered by the caveat [28–31]   1 thereto.) We showed that the mass of a firewall is exactly counterbalanced by the (negative) gravitational mass-energy accompanying its formation. Perhaps this may complement other lines of reasoning [4] disputing massiveness [3] of firewalls. In Section 7, we showed that equilibrium relativistic gravitational temperature gradients cannot be exploited to violate the Second Law of Thermodynamics. Auxiliary topics are discussed in the Notes.

Funding

There is no funding for this paper.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

There are no data that are not included in this paper.

Acknowledgments

I am very deeply grateful to Dr. Wolfgang Rindler, to Dr. Donald H. Kobe, and to Dr. James H. Cooke for lectures, seminars, discussions, and communications that greatly enhanced my understanding of relativity, and also of thermodynamics (especially relativistic thermodynamics). I am also very deeply grateful to Dr. Oliver Tennert and to Dr. L. Neslušan for insightful suggestions at ResearchGate concerning the clock paradox discussed in Note 1. Additionally, I thank Dr. Stan Czamanski and Dr. S. Mort Zimmerman for very interesting general scientific discussions over many years, Dr. Kurt W. Hess, Dan Zimmerman, and Robert H. Shelton for such discussions at times, and also Robert H. Shelton for very helpful advice concerning diction.

Conflicts of Interest

The author declares no conflicts of interest.

Notes

  1 Based on Birkhoff’s Theorem (see Refs. [20–28]), it is usually averred that in General Relativity—as in Newtonian gravitational theory—the gravitational field vanishes and the gravitational potential is negative and constant within an evacuated non-rotating spherical shell. This implies that spacetime is Minkowskian within the shell. (See, for example, Ref. [13], Section 12.2B.) However, there is a caveat [28–31]: To the contrary, it has also been averred that, in General Relativity—unlike in Newtonian gravitational theory—the gravitational field does not vanish within an evacuated non-rotating spherical shell, but instead that the field within the shell is directed radially outwards from the center [28–31]. This implies that the gravitational potential within the shell is negative but not constant, being least strongly negative at the center of the shell and most strongly negative at the inner surface of the shell. Moreover, contrary to the corresponding usual inference based Birkhoff’s Theorem (see Ref. [13], Section 11.2B and Refs. [20–28]), this further implies that spacetime is not Minkowskian within the shell [28–31]. (The discussion of this caveat in Ref. [28] is in the section entitled “Inside Spherical Shell” under the Talk tab thereof, and is intermediate in viewpoint between the standard interpretation of Birkhoff’s Theorem as per Refs. [20–27] and otherwise in Ref. [28] on the one hand, and that as per Refs. [29–31] on the other.) This at least helps to resolve a clock paradox in General Relativity [31]: If the gravitational field vanishes, the gravitational potential is negative and constant, and hence spacetime is Minkowskian within an evacuated non-rotating spherical shell, how is a clock within the shell to know that it is within the shell and thus at a negative gravitational potential, and hence that it must tick more slowly than a clock at r / r shell and hence at zero gravitational potential? For, like a clock at r / r shell , it would then see zero gravitational field and hence Minkowski spacetime. And according to General Relativity, a clock, like any other entity, interacts locally with a gravitational fieldno action at a distance. A non-vanishing gravitational field within an evacuated non-rotating spherical shell, which a clock therein can interact with locally, thus at least helps to resolve this clock paradox [28–31]: via local interaction with a non-vanishing gravitational field a clock at the center of the shell knows that it must tick more slowly than a clock at r / r shell [28–31], and a clock at the inner surface of the shell knows that it must tick more slowly yet [28–31]. Nonetheless a non-vanishing gravitational field within an evacuated non-rotating spherical shell does not alter any other inferences based on Birkhoff’s Theorem. If, on the contrary, the gravitational field does vanish, the gravitational potential is negative and constant, and hence spacetime is Minkowskian within an evacuated non-rotating spherical shell, then resolution of this clock paradox would seem to require either (i) local interaction of the clock’s gravitational field—which extends beyond the shell—with the shell’s gravitational field somehow being communicated to the clock itself [30] or (ii) local interaction of the clock with the shell’s gravitational potential [31]. The Aharonov-Bohm-effect counterpart of Option (i) is interpreting the Aharonov-Bohm effect as due to local interaction of an electron’s magnetic field with the magnetic field within a tightly-wound solenoid—the electron’s magnetic field penetrates into the solenoid—even though the electron itself sees only the solenoid’s magnetic vector potential and not the solenoid’s magnetic field [31]. (The electron must be moving relative to the solenoid in order for the Aharonov-Bohm-effect to occur and hence must generate a magnetic field in the reference frame of the solenoid. If the solenoid is tightly wound, the electron’s electric field cannot penetrate into it.) The Aharonov-Bohm-effect counterpart of Option (ii) is the standard interpretation of the Aharonov-Bohm effect: local interaction of the electron with the solenoid’s magnetic vector potential, which does not vanish outside of the solenoid [31]. (Although not related to the topics discussed in this paper, perhaps as a brief aside it should be noted that the Aharonov-Bohm effect is important in both theoretical and experimental investigations of electromagnetic quantum phenomena. See, for example, Imry, Y. In: Fraser, G., editor, Ref. [15]; Chapter 12 (especially Section 12.4–12.7).
  2 Note, however, that while the angular term of the Schwarzschild metric [the last term in all three lines of Equation (1)] is of identical Euclidean form at all  r 0 , by contrast radial ruler distance is d l = r S r 1 1 / 2 d t at r < r S , as opposed to d l = 1 r S r 1 / 2 d r at r r S . This obtains because the d t and d r terms of Schwarzschild metric [Equation (1)] switch sign as r S is crossed. See Ref. [13], Section 11.1 and 12.C–12.1E (especially Section 12.1D and 12.1E). Yet also note that in the line immediately following Equation (12.15) in Section 12.1E: At r < r S : r is referred to as a `time’—quotation marks in the original text—recognizing that while r is timelike, r is not time itself.
  3 The special case discussed in Ref. [34]—the excess (extra-Euclidean) vertical radial ruler distance of G M / 3 c 2 from the center to the surface of a non-rotating sphere of mass M and uniform density (in the weak-field limit, i.e., M r S c 2 / 2 G r r S = 2 G M / c 2 )—may help to clarify the vertical stretching of space from the Euclidean by gravity in general. It is a special case of the more general result discussed in Section 11.5 of Ref. [13]. By vertical it is of course meant perpendicular to the equipotential surface. The vertical direction does not in general coincide with the geometric center of a gravitator [see Ref. [13], Section 9.6 (especially the last two paragraphs)], but it does so coincide in the special case of a non-rotating spherical gravitator whose density varies at most only radially.
  4 English translations of Ref. [35] are provided in Refs. [36–38]. See also the Editor’s Note (Ref. [39]) and Ref. [40], which synopsize and discuss Ref. [35].
  5 Even if the classical vacuum might be construed as nothingness, the quantum-mechanical vacuum—space as it actually exists—certainly cannot. (See Ref. [13], pp. 418–419 and 480, Section 21.4, and Chapters 43–44; and Refs. [15,16].) If gravity stretches space, can space sustain tension? Since a medium capable of sustaining tension is required for the transmission of transverse waves [by contrast, longitudinal waves, e.g., sound, can travel through any (material, i.e., non-vacuum) medium], and since electromagnetic radiation is comprised of transverse waves, might space be construed as a latter-20th-century and 21st-century interpretation of the ether [sometimes spelled aether (the a is silent)] postulated in 19th-century physics? [See Ref. [13], Chapter 1 (especially Section 1.6–1.10); Ref. [40], Chapter 1 (especially pp. 8–20), and p. 66; and Ref. [36], pp. 495–496.] The conventional viewpoint is, of course, that electromagnetic waves serve as their own medium—their own ether—via the continual handoff of energy from transverse electric field to transverse magnetic field to transverse electric field ... . See Ref. [41], pp. 450–458 (especially pp. 452–453).
  6 By downwards it is of course meant perpendicular to the equipotential surface and towards a gravitator. Downwards is not in general towards the geometric center of a gravitator [see Ref. [13], Section 9.6 (especially the last two paragraphs)], but it is so in the special case of a non-rotating spherical gravitator whose density varies at most only radially. (Of course, upwards is in the opposite direction, i.e., perpendicular to the equipotential surface and away from the gravitator.)
  7 Because matter is not a continuum but is comprised of atoms, our opaque thermally-insulating spherical shell cannot be arbitrarily thin and therefore cannot have an arbitrarily small surface mass density ρ shell . Even to exist at all, it must be at least one atom thick. To be thermally-insulating, it must be opaque, and to be opaque it must be many atoms thick. (Opacity is a necessary but not sufficient condition for thermal insulation.) Hence (ignoring our speculations as per the last paragraph of Section 6) our spherical shell’s M shell / r shell ratio must be within a finite upper limit if it is not to be a black hole itself and suffer gravitational collapse: we must require the inequality M shell / r shell = 4 π ρ shell r shell 2 / r shell = 4 π ρ shell r shell < c 2 / 2 G r shell < c 2 / 8 π G ρ shell . But it certainly is feasible for r shell to greatly exceed λ r Wien , max = 15.90231998 r S [see the paragraph containing Equation (35)] while still meeting this inequality and hence without risk of the shell’s gravitational collapse: the strong inequality λ r Wien , max r shell < c 2 / 8 π G ρ shell , indeed, even the double strong inequality λ r Wien , max r shell c 2 / 8 π G ρ shell , is very easily met. This is sufficient is for r shell to effectively obtain for all practical purposes.
  8 Equations (2)–(6), (9), and (10) are in accordance with considerations of Unruh radiation and the equivalence principle (see Refs. [68,69]). An object undergoing acceleration a in Minkowski spacetime experiences Unruh radiation at temperature T U = a 2 π c k . Force f = m c 4 4 G M 1 r S r 1 / 2 = G M m r S 2 1 r S r 1 / 2 is required to dangle a mass m at r S from a higher altitude r > r S (with a massless string) above a Schwarzschild black hole of mass M: see Ref. [13], Section 12.2 [especially Equation (12.17)]. The corresponding acceleration is a = f m = c 4 4 G M 1 r S r 1 / 2 = G M r S 2 1 r S r 1 / 2 and hence the corresponding Unruh-radiation temperature is T U , r = a 2 π c k = 2 π c k × c 4 4 G M 1 r S r 1 / 2 = 1 r S r 1 / 2 c 3 8 π G k M = 1 r S r 1 / 2 T H , r = T H , r . In accordance with the equivalence principle, T U , r is equal to T H , r as per Equation (3). In the limit r , in accordance with the equivalence principle, T U , r = c 3 8 π G k M = T H , r as per Equation (2). {See Ref. [13], Section 12.2 [especially Equation (12.17)] and Section 12.6.} But a caveat: It is important to note that: Hawking-radiation temperature T H is by Equations (2) and (3) a function of the gravitational potential  Φ . By contrast, Unruh-radiation temperature T U is a function of the motional acceleration a in Minkowski spacetime, so prima facie the equivalence principle might seem to suggest that it be the same function of the magnitude G = d Φ / d l of the gravitational acceleration, i.e., the same function of the magnitude of the gradient of the potential rather than a function of the potential itself (whether of a black hole or a non-black hole). But this incorrectly implies that T U need not in general be equal to T H : e.g., at large enough r away from a Schwarzschild black hole or for a sufficiently weak Schwarzschild non-black hole that the Newtonian approximation a = f / m = G M / r 2 is valid with negligible error, this incorrectly implies that T U , r = a 2 π c k = G M 2 π c k r 2 , and in the limit r , that T U , r = 0 —in disagreement with T H , r > 0 as per Equation (2), T H , r as per Equation (3), and Tolman’s [8,9] generalization, bolstered with Ehrenfest [10] (see also Garrod [11]), as per Equations (4)–(6). The correct correlation between T H and T U is that obtained as per Ref. [13], Section 12.2, especially the paragraph containing Equation (12.17): via dangling a mass at r S from a higher altitude r r S (with a massless string). With respect to the correct correlation between T T and T U , perhaps our plausibility argument as per Equations (7)–(10) and the associated discussions may at least serve as a starting point.
  9 Usually it is assumed that electromagnetic radiation can be tracked to a source, but Maxwell’s equations do not require this. This was pointed out to me by Dr. James H. Cooke in a private communication in the 1980s.
  10 The concepts of microscopic reversibility and detailed balance require modifications in cases of (i) time-symmetry-violating dynamics and (ii) collisions between unsymmetrical molecules even given non-time-symmetry-violating dynamics. See, for example, Ref. [95] concerning (i) and Ref. [96] concerning (ii). But these modifications do not apply with respect to electromagnetic radiation in general and hence with respect to equilibrium blackbody radiation in particular. Hence the analyses provided in Refs. [8–11,64,65] concerning equilibrium blackbody radiation are completely valid.
  11 Auxiliary phenomena that might have contributed to the end of the dinosaurs’ reign could have included a surge in volcanic activity, the impact of a secondary asteroid if the primary impactor had a satellite, etc.
  12 Reference [104] states that owing to quantum gravitational corrections, Hawking radiation is not exactly Planckian, i.e., not exactly blackbody, and thus not exactly maximum-entropy and hence a carrier of information. But this assumes that a black hole radiates into empty space. What happens if, instead, a black hole is enclosed concentrically by an opaque thermally-insulating spherical shell? Initially upon emission from the black hole, Hawking radiation emanating from the black hole would still carry information. But the Hawking radiation emanating from the black hole would then be thermalized to an exactly Planckian distribution within the spherical shell. Would its information then be lost? Or would the information be preserved, even if only in latent form, even after thermalization? Also, a few caveats concerning Ref. [104] are quoted in Ref. [105]. References [106] and [107] seem especially pertinent with respect to the last paragraph of Section 6, because they discuss the possibility—at least in principle, even if not in practice with realizable technology—of experimentally determining whether the black-hole information paradox is resolved via firewalls, as we discussed qualitatively in the last paragraph of Section 6, or via complementarity, according to which the interior of a black hole and Hawking radiation are not independent, but correlated.

References

  1. Polchinski, J. Burning Rings of Fire. Scientific American Special Edition “Black Holes”. Winter 2021, 30 (1), 58–63.
  2. Giddings, S.B. Escape from a Black Hole. Scientific American Special Edition “Black Holes”. Winter 2021, 30 (1), 70–77.
  3. Abramowicz, M.A.; Kluzniak, W.; Lasota, J.P. Mass of a Black Hole Firewall. Phys. Rev. Lett. 2014, 112, Article Number 091301, 4 pages. See also references cited therein.
  4. Kaplan, D.E.; Rajendran, S. Firewalls in general relativity. Phys. Rev. D 2019, 99 (4), Article Number 044033, 8 pages. See also references cited therein.
  5. Harlow, D. Jerusalem lectures on black holes and quantum information. Reviews of Modern Physics 2016, 88, Article Number 015002, 58 pages. See also references cited therein.
  6. Dai, X. The Black Hole Paradoxes and Possible Solutions. J. Phys. Conf. Ser. 2020, 1634, Article Number 012088, 7 pages. See also references cited therein.
  7. Firewall (physics). Wikipedia. https://en.wikipedia.org/wiki/Firewall_(physics) (2023). See also references cited therein. (accessed on 4 March 2024).
  8. Tolman, R.C. On the Weight of Heat and Thermal Equilibrium in General Relativity. Phys. Rev. 1930, 35, 904–924. [Google Scholar] [CrossRef]
  9. Tolman, R.C.; Ehrenfest, P. Temperature Equilibrium in a Static Gravitational Field. Phys. Rev. 1930, 36, 1791–1798. [Google Scholar] [CrossRef]
  10. Tolman, R.C. Relativity, Thermodynamics, and Cosmology; Oxford University Press: Oxford, UK, 1934; unabridged and unaltered republication by Dover, New York, NY, USA, 1987; Sections 128 and 129.
  11. Garrod, C. Statistical Mechanics and Thermodynamics; Oxford University Press: New York, NY, USA, 1995; Exercises 7.29 and 7.30. [Google Scholar]
  12. Mann, A. Disappearing Act. Sci. Amer. September 2023, 329(2), 8–10. [Google Scholar]
  13. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Sections 9.3–9.4, 11.1–11.5, and 12.2, the last paragraph on p. 326 and the first paragraph on p. 327, pp. 367, 374, 384–385, and Exercise 3.27. [Google Scholar]
  14. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation, First Princeton (Kindle) edition; Princeton University Press: Princeton, NJ, USA, 2017; pp. 12–13, in Box 17.2 Items 2’c and 2’d on p. 419, and Item 6 on pp. 426–428, the last paragraph of Section 20.6 on p. 480, Chapter 43 (especially Sections 43.1 and 43.4), and Chapter 44 (especially Sections 44.3 and 44.6).
  15. Adler, R. Gravity. In: Fraser, G., editor. The New Physics for the Twenty-First Century, First paperback edition; Cambridge University Press: Cambridge, UK, 2009; Chapter 2. (See especially Sections 2.9 and 2.10.).
  16. Green, M.B. Superstring Theory. In: Fraser, G., editor. The New Physics for the Twenty-First Century, First paperback edition; Cambridge University Press: Cambridge, UK, 2009; Chapter 5. (See especially Sections 5.3 and 5.4.).
  17. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006, 2017; Section 1.16, Chapter 9 (especially Sections 9.3 and 9.4), and Sections 11.2A, 11.6, and 12.2 (especially Section 12.2).
  18. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation, First Princeton (Kindle) edition; Princeton University Press: Princeton, NJ, USA, Sections 7.2–7.4.
  19. CODATA Internationally recommended 2018 values of the Fundamental Physical Constants. National Institute of Standards and Technology (NIST). https://physics.nist.gov/cuu/Constants/ (2018). See the category entitled “Universal constants”. (accessed on 4 March 2024).
  20. Tolman, R.C. Relativity, Thermodynamics, and Cosmology; Oxford University Press: Oxford, UK, 1934; unabridged and unaltered republication by Dover, New York, NY, USA, 1987; Section 99.
  21. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006 2017; Sections 11.2B, 12.2 (especially the last three paragraphs), 14.4, 16.1H, and 17.4, and Exercise 12.12.
  22. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation, First Princeton (Kindle) edition; Princeton University Press: Princeton, NJ, USA, 2017; Section 32.2, Exercise 32.1, and Section 32.3.
  23. Ohanian, H.C.; Ruffini, R. Gravitation and Spacetime, 3rd (Kindle) edition. Cambridge University Press, Cambridge, UK, 2013; the 6th- through 3rd-to-last paragraphs of Section 7.4, the second paragraph preceding that containing Equation (8.14), the paragraph containing Equations (9.21)–(9.25) and the immediately following paragraph, and Appendix A.5.
  24. Carroll, B.W.; Ostlie D.A. An Introduction to Modern Astrophysics, 2nd (Kindle) edition. Cambridge University Press, Cambridge, UK, 2017; the section entitled “Extending Our Simple Model to Include Pressure” on pp. 1160–1162, Problem 10.2, and the section entitled “The Development of Adiabatic Density Fluctuations” on pp. 1248–1251.
  25. Jebsen, J.T. Über die allgemeinen kugelsymmetrischen Lösungen der Einsteinschen Gravitationsgleichungen im Vakuum. Arkiv för Matematik, Astronomi och Fysik 1921,15 (18), 9 pages.
  26. Deser, S. Introduction to Jebsen’s paper. Gen. Relativ. Gravit. 2005, 37, 2251–2252. [Google Scholar] [CrossRef]
  27. Jebsen, J.T. (translation from the German original by Antonci, S.; Liebscher, D.). On the general spherically symmetric solutions of Einstein’s gravitational equations in vacuuo. Gen. Relativ. Grav. 2005, 37, 2253–2259. [Google Scholar] [CrossRef]
  28. Birkhoff’s theorem (relativity). Wikipedia. https://en.wikipedia.org/wiki/Birkhoff%27s_theorem_(relativity) (2023). See the references cited therein, and the caveat discussed in the section entitled “Inside Spherical Shell” under the Talk tab. (accessed on 4 March 2024).
  29. Neslušan, L. The second rise of general relativity in astrophysics. Modern Physics Letters A 2019, 34 (30), Article Number 1950244, 24 pages.
  30. deLyra, J.L.; Carneiro, C.E.I. Complete Solution of the Einstein Field Equation for a Spherical Distribution of Polytropic Matter. 22 pages. Available online at publica-sbi.if.usp.br/PDFs/pd1731.pdf and portal.if.usp.br/bib/sites/portal.if.usp.br.bib/files/PDFs/pd1731.pdf (accessed on 4 March 2024).
  31. A spherical-shell clock paradox in General Relativity? Available online at https://www.researchgate.net/post/a_spherical-shell_clock_paradox_in_General_Relativity (accessed on 4 March 2024).
  32. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Chapters 11 and 12. [Google Scholar]
  33. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Chapters 9 and 11 (in Chapter 11 see especially Sections 11.1–11.6). [Google Scholar]
  34. Feynman, R.P.; Leighton, R.B.; Sands, M. The Feynman Lectures on Physics, Kindle edition; University of California Press: Berkeley, CA, USA, 2015; Volume II, Chapter 42 (especially Sections 42-2, 42-3, and 42-9). [Google Scholar]
  35. Sakharov, A.D. Vacuum Quantum Fluctuations in Curved Space and the Theory of Gravitation. Doklady Akademii Nauk SSSR 1967, 177, 70–71. [Google Scholar]
  36. Sakharov, A.D. Vacuum Quantum Fluctuations in Curved Space and the Theory of Gravitation. Soviet Physics Doklady 1968, 12, 1040–1041. [Google Scholar]
  37. Sakharov, A.D. Vacuum Quantum Fluctuations in Curved Space and the Theory of Gravitation. Soviet Physics Uspekhi 1991, 34, 394. [Google Scholar] [CrossRef]
  38. Sakharov, A.D. Vacuum Quantum Fluctuations in Curved Space and the Theory of Gravitation. Gen. Relativ. Gravitat. 2000, 32, 365–367. [Google Scholar] [CrossRef]
  39. Schmidt, Hans-Jürgen. Editor’s Note: Vacuum Quantum Fluctuations in Curved Space and the Theory of Gravitation. Gen. Relativ. Gravitat. 2000, 32, 361–363. [CrossRef]
  40. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation, First Princeton (Kindle) edition; Princeton University Press: Princeton, NJ, USA, 2017; pp. 426–428 and 1206–1208. [Google Scholar]
  41. Epstein, L.C. Relativity Visualized; Insight Press: San Francisco, CA, USA, 1997; Chapter 1, (especially pp. 8–20), and p. 66. [Google Scholar]
  42. Epstein, L.C. Thinking Physics: Understandable Practical Reality, 3rd edition; Insight Press: San Francisco, CA, USA, 2015; pp. 495–496. [Google Scholar]
  43. . Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Sections 1.12–1.15, 11.2A, 12.2, and 12.6 (especially Section 12.6). [Google Scholar]
  44. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation, First Princeton (Kindle) edition; Princeton University Press: Princeton, NJ, USA, 2017; Item 1.B.d on p. xxxvi and Item 4.B.f on p. xxxviii. [Google Scholar]
  45. Adler, R. Gravity. In Fraser, G., editor. The New Physics for the Twenty-First Century, First paperback edition; Cambridge University Press: Cambridge, UK, 2009; Chapter 2. (See especially Section 2.10.). [Google Scholar]
  46. Ohanian, H.C.; Ruffini, R. Gravitation and Spacetime, 3rd (Kindle) edition. Cambridge University Press, Cambridge, UK, 2013; Section 8.7.
  47. Carroll, B.W.; Ostlie, D.A. An Introduction to Modern Astrophysics, 2nd (Kindle) edition. Cambridge University Press, Cambridge, UK, 2017; pp. 644–646 and Problem 17.23.
  48. Scully, M.; Sokolov, A.; Svidzinsky, A. Virtual Photons: From the Lamb Shift to Black Holes, Optics & Photonics News, February 2018, 34–40.
  49. Hawking radiation. Wikipedia. https://en.wikipedia.org/wiki/Hawking_radiation (2023). See also references cited therein. (accessed on 4 March 2024).
  50. Lang, K.R. Essential Astrophysics, Kindle edition; Springer-Verlag: Berlin, Germany, 2013; Sections 2.1–2.4 (for background material), Sections 15.1 and 15.2, and p. 608. [Google Scholar]
  51. García-Bellido, S.; Clesse, S. Black Holes from the Beginning of Time. Scientific American Special Edition “Black Holes” Winter 2021, 30 (1), 26–31.
  52. Carr, B.; Kühnel, F. Primordial Black Holes as Dark Matter Candidates: Recent Developments. Annual Review of Nuclear and Particle Science 2020, 70, 355–394, See also references cited therein. [Google Scholar] [CrossRef]
  53. Villanueva-Domingo, P.; Mena, O.; Palomarez-Ruiz, S. A Brief Review on Primordial Black Holes as Dark Matter, Front. Astron. Space Sci. 2021, 8. Article Number 681084, 10 pages. See also references cited therein. [Google Scholar]
  54. Green, A.M.; Kavanagh, B.J. Primordial black holes as a dark matter candidate, J. Phys. G. Nucl. Part. Phys. 2021, 48. Article Number 043001, 29 pages. See also references cited therein. [Google Scholar] [CrossRef]
  55. Kainulainen, K.; Nurmi, S.; Schaippacasse, E.D.; Yanagida, T.T. Can primordial black holes as all dark matter explain fast radio bursts? Phys. Rev. D 2021, 104. Article Number 123033, 7 pages. See also references cited therein. [Google Scholar] [CrossRef]
  56. Stuart, C. The First Black Holes, Science Focus November 3, 2021, 70–77.
  57. Carr, B. Black holes from a previous universe. New Scientist April 1–7, 2023,257 (3432), 46–49.
  58. Fumagalli, J.; Renaux-Petel, S.; Ronayne, J.W.; Witkowski, L.T. Turning in the landscape: A new mechanism for generating primordial black holes. Physics Letters B 2023, 841. Article Number 137921, 8 pages. [Google Scholar] [CrossRef]
  59. Sutter, P. Tiny primordial black holes could have created their own Big Bang. https://www.space.com/primordial-black-holes-create-big-bang (accessed on 4 March 2024).
  60. Dienes, K.R.; Heurtier, L.; Huang, F.; Kim, D.; Tait, T.M.P.; Thomas, B. Primordial Black Holes Place the Universe in Statis. arXiv: 2212.01369v2 [astro-ph.CO] 28 March 2023, 25 pages. (accessed on 4 March 2024).
  61. Primordial black hole. Wikipedia. https://en.wikipedia.org/wiki/Primordial_black_hole (2023). See also references cited therein. (accessed on 4 March 2024.
  62. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Section 12.2, especially the paragraph containing Equation (12.17). [Google Scholar]
  63. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Section 1.16, Chapter 9 (especially Sections 9.3, 9.4, 9.6, and 9.7), and Sections 11.1–11.6, 11.13, 12.1, and 12.2 (especially Sections 11.1, 11.2, and 12.2). [Google Scholar]
  64. Reif, F. Fundamentals of Statistical and Thermal Physics, Kindle edition; Waveland Press: Long Grove, IL, 2009; Sections 9.13–9.15 and Problems 9.9–9.13. [Google Scholar]
  65. Baierlein, R. Thermal Physics; Cambridge University Press: Cambridge, UK, 1999; introductory material for Chapter 6 on p. 116, Section 6.1, Items 1–7 in Section 6.6, and Problems 1–15 for Chapter 6. [Google Scholar]
  66. CODATA Internationally recommended 2018 values of the Fundamental Physical Constants. National Institute of Standards and Technology (NIST). https://physics.nist.gov/cuu/Constants/ (2018). See the category entitled “Physico-chemical constants”. (accessed on 4 March 2024).
  67. Williams, D.R. Sun Fact Sheet. National Space Science Data Center 2022. https://nssdc.gsfc.nasa.gov/planetary/factsheet/sunfact.html (accessed on 4 March 2024).
  68. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation, First Princeton (Kindle) edition; Princeton University Press: Princeton, NJ, USA, 2017; Item C of Box 24.1, Figure 24.1, and the caption in the title of Chapter 33. [Google Scholar]
  69. Wheeler, J.A.; Ford, K. Geons, Black Holes, and Quantum Foam: A Life in Physics; Kindle edition: W. W. Norton, New York, NY, USA, 2000; Chapters 10 and 13. [Google Scholar]
  70. Lang, K.R. Essential Astrophysics, Kindle edition; Springer-Verlag: Berlin, Germany, 2013; Sections 13.7.1 and 13.8.1. [Google Scholar]
  71. Carroll, B.W.; Ostlie D.A. An Introduction to Modern Astrophysics, 2nd (Kindle) edition. Cambridge University Press, Cambridge, UK, 2017; pp. 633–634 and Problem 17.10.
  72. Black hole. Wikipedia. https://en.wikipedia.org/wiki/Black_hole (2023), especially the second paragraph and the section entitled “History”. See also references cited therein. (accessed on 4 March 2024).
  73. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Sections 1.12–1.15, Exercise 1.7, and Sections 8.4 and 14.1. [Google Scholar]
  74. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation, First Princeton (Kindle) edition; Princeton University Press: Princeton, NJ, USA, 2017; Sections 7.4 and 10.1, Chapter 16, and Section 38.6. [Google Scholar]
  75. Fritz, S. Solar Radiant Energy and Its Modification By the Earth and Its Atmosphere. In Malone, T.F., editor. Compendium of Meteorology; American Meteorological Society: Boston, MA, USA, 1951; pp. 13–33, (See especially pp. 13–19.). [Google Scholar]
  76. Wallace, J.M.; Hobbs, P.V. Atmospheric Science, 2nd edition; Elsevier: Amsterdam, The Netherlands, 2006; Section 4.3 (especially Subsections 4.3.1–4.3.3). [Google Scholar]
  77. Houghton, H.G. Physical Meteorology; MIT Press: Cambridge, MA, USA, 1985; Chapter 3 (especially Sections 3.3 and 3.4). [Google Scholar]
  78. Sunlight. Wikipedia. https://en.wikipedia.org/wiki/Sunlight 2023. See also references cited therein. (accessed on 4 March 2024).
  79. Duley, W.W. Blackbody Radiation in Small Cavities. Am. Jour. Phys. 1972, 40, 1337–1338. [Google Scholar] [CrossRef]
  80. Baltes, H.P. Comment on Blackbody Radiation in Small Cavities. Amer. Jour. Phys. 1974, 42, 505–507. [Google Scholar] [CrossRef]
  81. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation, First Princeton (Kindle) edition; Princeton University Press: Princeton, NJ, USA, 2017; Exercises 22.15–22.17 (especially Exercise 22.17) on pp. 588–589, and Box 29.2. [Google Scholar]
  82. Wien’s displacement law. Wikipedia. https://en.wikipedia.org/wiki/Wien%_27s_displacement law (2023). See also references cited therein. (accessed on 4 March 2024).
  83. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Sections 1.16, 11.2, and 12.2. [Google Scholar]
  84. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation, First Princeton (Kindle) edition; Princeton University Press: Princeton, NJ, USA, 2017; Section 7.2. [Google Scholar]
  85. Rindler, W. Introduction to Special Relativity, 2nd (Kindle) edition; Oxford University Press: New York, NY, USA, 1991; Section 33. [Google Scholar]
  86. Aharoni, J. The Special Theory of Relativity, 2nd edition; Oxford University Press: Oxford, UK, 1965; unabridged and unaltered republication by Dover, New York, NY, USA, 1985; pp. 158–160 and 324. [Google Scholar]
  87. Wheeler, J.A.; Ford, K. Geons, Black Holes, and Quantum Foam: A Life in Physics; Kindle edition; W. W. Norton, New York, NY, USA, 2000; Chapters 10 and 11 (especially pp. 234–239, 252–253, and 256–257). [Google Scholar]
  88. Klimets, A.P. On the fundamental role of massless form of matter in physics 2021, 33 pages. (See especially Chapter 3.) https://philpapers.org/archive/ALXOTF.pdf (accessed on 4 March 2024).
  89. Geon (physics). Wikipedia. https://en.wikipedia.org/wiki/Geon_(physics) (2023). See also references cited therein. (accessed on 4 March 2024).
  90. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Section 11.5. [Google Scholar]
  91. Tolman, R.C. Relativity, Thermodynamics, and Cosmology; Oxford University Press: Oxford, UK, 1934; unabridged and unaltered republication by Dover, New York, NY, USA, 1987; Sections 82 and 94–96. [Google Scholar]
  92. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Chapter 11 (especially Sections 11.1–11.6), and Sections 12.1 and 12.2. [Google Scholar]
  93. Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation, First Princeton (Kindle) edition; Princeton University Press: Princeton, NJ, USA, 2017; Part VII (especially Chapter 31). [Google Scholar]
  94. Rindler, W. Relativity: Special, General, and Cosmological, 2nd (Kindle) edition; Oxford University Press: Oxford, UK, 2006; Section 7.2.F., Chapter 9 (especially Sections 9.3, 9.4, 9.6, and 9.7), and Sections 11.1, 11.2, 11.5, 11.6, and 12.2. [Google Scholar]
  95. Reif, F. Fundamentals of Statistical and Thermal Physics, Kindle edition; Waveland Press: Long Grove, IL, 2009; Sections 9.15, 15.1, 15.2, and 15.18 (in Section 15.18 see especially pp. 598–600); also, Problem 9.14. [Google Scholar]
  96. Davies, P.C.W. The Physics of Time Asymmetry, 1977 edition; University of California Press: Berkeley, CA, USA, 1977; Section 6.4. [Google Scholar]
  97. Tolman, R.C. The Principles of Statistical Mechanics; Oxford University Press, Oxford, UK, 1938; unabridged and unaltered republication by Dover, New York, NY, USA, 1979; Sections 40–43.
  98. Carroll, B.W.; Ostlie D.A. An Introduction to Modern Astrophysics, 2nd (Kindle) edition. Cambridge University Press, Cambridge, UK, 2017, pp. 842–844.
  99. Roncaglia, M. On the Conservation of Information in Quantum Physics. Found. Phys. 49, 1278–1286 (2019).
  100. Strauch, F.W. Resource Letter QI-1: Quantum Information. Amer. Jour. Phys. 84 (7), 495–507 (2016). (See especially Section VII.).
  101. Giddings, S.B. Black Holes in the Quantum Universe. Phil. Trans. Royal Soc. A 2019, 377, Article Number 20190029, 13 pages.
  102. Scientific American Special Report: Black Hole Mysteries Solved. Sci. Amer. September 2022, 327 (3), 28–53: (i) Moskowitz, C. Black Hole Mysteries Solved. Sci. Amer. September 2022, 327 (3), 28–29. (ii) Musser, G. Paradox Resolved. Sci. Amer. September 2022, 327 (3) 30–33. (iii) Almheiri, A. Black Holes, Wormholes, and Entanglement. Sci. Amer. September 2022, 327 (3), 34–41. (iv) Shaghoulian, E. A Tale of Two Horizons. Sci. Amer. September 2022, 327 (3), 42–47. (v) Fletcher, S. Portrait of a Black Hole. Sci. Amer. September 2022, 327 (3), 48–53. See also the references cited in From Our Archives on p. 53 [in (v)]. The black-hole information paradox, and its resolution, is discussed mainly in (i)–(iv).
  103. Ananthaswamy, A. The Holographic Universe Turns 25. Sci. Amer. March 2023, 328(3), 58–61. [Google Scholar]
  104. Calmet, X.; Hsu, S.D.H.; Sebastianutti, M. Quantum gravitational corrections to particle creation by black holes. Phys. Lett. B 2023. Article Number 137820. [Google Scholar] [CrossRef]
  105. Crane, L. Information can survive a black hole. New Scientist April 1–7, 2023, 258 (3433), 8.
  106. Wilkins, A. A way to solve the black hole paradox. New Scientist January 13–19, 2024, 261 (3473), 10–11.
  107. Bousso, R.; Penington, G. Islands Far Outside the Horizon. arXiv:2312.03078v3 [hep-th] 20 December 2023, 28 pages.
  108. Black hole information paradox. Wikipedia. https://en.wikipedia.org/wiki/Black_hole_information_paradox (2023). See also references cited therein. (accessed on 4 March 2024).
  109. Reif, F. Fundamentals of Statistical and Thermal Physics, Kindle edition; Waveland Press: Long Grove, IL, 2009; Sections 2.3 and 6.1–6.4 (in Section 6.3 see especially the subsection entitled “Molecule in an ideal gas in the presence of gravity”). [Google Scholar]
  110. Wark Jr., K.; Richards, D.E. Thermodynamics, 6th edition; WCB/McGraw-Hill: Boston, MA, USA, 1999; p. 11 and Section 6-3-5. [Google Scholar]
  111. Wark Jr., K.; Richards, D.E. Thermodynamics, 6th edition; WCB/McGraw-Hill: Boston, MA, USA, 1999. [Google Scholar]
  112. Wallace, J.M.; Hobbs, P.V. Atmospheric Science, 2nd edition; Elsevier: Amsterdam, The Netherlands, 2006; Section 3.2. [Google Scholar]
  113. Holton, J.R.; Hakim, G.J. An Introduction to Dynamic Meteorology, 5th edition; Elsevier: (Amsterdam, The Netherlands, 2013; Section 1.4.1. [Google Scholar]
  114. Schroeder, D.V. Thermal Physics; Addison Wesley Longman: San Francisco, CA, USA, 2000; Section 1.1, and Problems 1.16, 3.37, and 6.14. [Google Scholar]
  115. Schroeder, D.V. Thermal Physics; Addison Wesley Longman: San Francisco, CA, USA, 2000; Problem 1.16. [Google Scholar]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

© 2024 MDPI (Basel, Switzerland) unless otherwise stated