2407.06207v1
2407.06207v1
Abstract
Starting from a new understanding of the vacuum energy problem based on the
combination of the phase space regularization and the holographic bound, we argue that
quantum gravity should be understood as gravitized quantum theory, that is, quantum
theory wherein the geometry and topology of the state-space if fully dynamical, in
analogy with the dynamical nature of spacetime in Einstein’s general relativity. Apart
from the vacuum energy problem viewed as a quantum gravity problem, we discuss
the “smoking gun” experiments involving higher order quantum interference, as well
as experimental probes of the statistics of spacetime quanta. Finally, we address the
conundrum of the intricately patterned spectrum of masses of elementary particles as
well as their mixing angles, as another telltale problem of quantum gravity viewed as
gravitized quantum theory.
∗
Prepared for the Proceedings of the 3rd Conference on Nonlinearity (September, 2023), held at the
Serbian Academy of Nonlinear Sciences, Belgrade, Serbia
†
[email protected]
‡
[email protected]
Contents
1 Introduction: Why Quantum Gravity? 1
References 51
1 Introduction: Why Quantum Gravity?
A final, comprehensive, cohesive and consistent unification of quantum physics and gravi-
tation (relativistic as well non-relativistic) — dubbed quantum gravity (QG) — has been
a coveted goal for the better part of the past century. Yet, even the very concept of what,
exactly, QG ought to be (in fact, even whether any such a thing can exist) continues to be
hotly debated (to say the least [1,2], and references therein), especially in the absence of any
guiding and illuminating sharp empirical facts.
Indicative of its profoundly vexing difficulty, over two dozen theoretical approaches to
QG (see [1, 2] and references therein) have been developed by now:1 the effective field the-
ory (EFT) approach, asymptotic safety, supergravity, string theory, holography (AdS/CFT
and generalizations: de Sitter (dS), celestial, corner symmetry), Euclidean quantum gravity,
topological (and categorical) quantum field theory, canonical quantum gravity, loop quantum
gravity (Hamiltonian/spin networks and covariant/spin foams), causal sets, quantum cos-
mology, group field theory, emergent gravity in condensed matter, Regge calculus, (causal)
dynamical triangulation, non-commutative geometry, twistor theory, Horava-Lifshitz gravity,
analog gravity models, modified/massive gravity, gauge theories of gravity, non-local theo-
ries of gravity, shape dynamics, entropic gravity, quantum gravity phenomenology, quantum
graphity, gravitized quantum theory, etc. There are also many key questions to be answered
by quantum gravity, which may be grouped as follows. Questions about spacetime: res-
olution of singularities (black hole and cosmological), quantum black holes (including black
hole (BH) entropy, information puzzle), astrophysics of the (resolved) BH singularity, cosmol-
ogy of the (resolved) initial singularity, fine tuning of the initial state, early universe cosmol-
ogy, relevance of quantum gravity for structure formation, quantum structure of spacetime
and its phenomenology, the vacuum energy problem, etc. Questions about matter: the
observed hierarchy of scales, origin of the Standard Models of particle physics and cosmology,
masses and couplings of fundamental particles, dark matter and dark energy, baryogenesis,
etc. Conceptual questions: emergence of quantum theory, quantum/classical transition,
quantum measurement problem, topology change, problem of time, closed timelike curves,
emergence of spacetime, origin of inertia, etc. A tall order indeed!
Herein, we consider this difficult problem rather literally from the ground up: starting
with the long-standing and oft-ridiculed as an impossible problem in effective field the-
ory (EFT), the vacuum energy, i.e., the cosmological constant, Λcc , which has been mea-
sured [3, 4], and which could be considered as one established empirical fact of quantum
gravity. In particular, we present a new computation of the vacuum energy fundamen-
tally based on the short-distance/long-distance interplay between the phase space geometry
1
Even in a review such as this, there can be no hope for any semblance of doing justice in reviewing these
approaches and important questions, so we list the most prominent among them to indicate the breadth of
the field, and to help the interested reader with starting their own searching efforts (see also, for example,
the “International Society for Quantum Gravity” YouTube channel, https://www.youtube.com/@isqg423,
and the lecture https://www.youtube.com/watch?v=7T6pinf7fLg ).
1
and the holographic bound. Motivated by this new calculation of the vacuum energy that
matches the observed value (interpreted as the cosmological constant in Einstein’s gravita-
tional equations, and in principle the first measured quantum gravity observable), we argue
that the essential ingredient of quantum gravity is a dynamical quantum phase space, and
thus a dynamical form of the Born rule used to define general quantum probabilities and
observables of quantum gravity. This identifies gravitization of quantum theory 2 as the key
feature by which the approach presented in this review differs from traditional approaches to
quantum gravity. We present a general discussion of quantum spacetime rewriting of quan-
tum theory and a dynamical quantum spacetime formulation found in a non-commutative
and phase-space-like formulation of string theory, the metastring [5]. The gravitization of
quantum theory sheds light on the origin of quantum theory and quantum field theory and
so gravitized quantum theory can be understood as a metaquantum theory. Also gravitized
quantum theory in turn implies very specific experimental probes of higher order quantum
interference processes which are impossible in local quantum field theory. Furthermore,
the statistics of the “atomic” constituents of spacetime is illuminated in this context, and
contrasted to the familiar spin-statistics relation established for the matter degrees of free-
dom. The statistics of spacetime quanta can be also empirically probed using gravitational
interferometry.
The calculation of the vacuum energy can be extended to compute the masses (gravi-
tational charges) of fundamental particles in the matter sector and find a cascade of seesaw-
determined mass-scales, which befit the Standard Model (SM) extraordinarily well. This
intricate pattern of SM fermion masses [6] and the vacuum energy problem [7] are the two
most vexing issues in fundamental physics — and so also of quantum gravity. Following [8,9]
and sharpening the recent analysis [10], we provide a conceptually more cohesive and coher-
ent framework and show that its unifying, ultraviolet (UV)/infrared (IR) mixing solution to
these two problems of fundamental physics also addresses the gauge hierarchy (Higgs mass)
problem as well as the problem of mixing both in the quark and neutrino sectors [10, 11].
In what follows we emphasize the quantum gravitational nature of these problems,
represented by the explicit appearance of widely separated short-distance and long-distance
scales. Thus quantum gravity is not a (purely) Planck scale phenomenon, as it finds its
physical manifestations at all scales via the spectrum of elementary particles. Furthermore,
low energy phenomena of higher order quantum interference as well as the experimental
probes of the statistics of spacetime quanta, are indicative of both the infrared and ultraviolet
nature of quantum gravity phenomenology. From our viewpoint, one of the most pressing
issues in the research in quantum gravity (where theory reigns supreme) is the development
of various experimental probes, at various scales, of such quantum gravity phenomenology.
This hallmark UV/IR-mixing (which evidently transcends EFT), and quantum contex-
tuality (in contradistinction to anthropic reasoning [12]) are key features of our solution to
2
By “gravitizing” we mean rendering the geometry and even topology of the quantum theory state-space
fully dynamical, akin to how spacetime becomes dynamical in Einstein’s theory of gravity.
2
the vacuum energy problem. It also involves the Born geometry [13] (a kind of generalized
mirror symmetry [14]) of the stringy (chiral phase-space-like) spacetime and its modular
polarization, the Bekenstein [15], i.e., holographic [16–18] bound, combined with stringy
modular invariance [19, 20], and a stringy formula for the Higgs mass [21]. These concepts
are in fact crucial to addressing the problem of SM fermion masses (gravitational charges)
and mixing matrices in the context of string theory/quantum gravity, and lead to obser-
vations and mechanisms that are all realized within string theory — the only consistent
perturbative theory of both quantum gravity and SM-like matter. They give rise to the con-
crete results summarized in Table 1 (depicted in Figure 1), which correlate very well with
Bjorken’s recent work [22, 23]. Our analysis of their correlations is what underlies a formu-
lation of what quantum gravity ought to be — a gravitized quantum theory, as announced
in the title.
This review builds on our previous paper [10] (especially §§ 2 and 5) and is organized
as follows: Key aspects of the cosmological constant (cc) problem are discussed in § 2, which
we review (§ 2.1) first in point-particle quantum field theory (QFT), and then also in string
theory. Its recent resolution [8] (see also [9]) is then presented in § 2.2, exhibiting that it
(A) connects quantized phase space properties with the Bekenstein (holographic) bound in
a gravitational setting [16], and (B) is realized both in QFT (§ 2.3), as well as in string
theory (§ 3.5). Given the computation of the vacuum energy based on (in general) dynam-
ical phase space, we address the general argument for the formulation of quantum gravity
as gravitized quantum theory in § 3. Here we usher the reader from the concepts of mod-
ular spacetime as a quantum model of spacetime and the associated Born geometry to the
metastring formulation of the general non-commutative and T-duality covariant string theory
in dynamical quantum spacetime and dynamical Born geometry. Viewing quantum grav-
ity as gravitized quantum theory also sheds light on the origins of quantum field theory and
quantum mechanics, both based on the Born rule (Bornian quantum theory). Distinguishing
experimental probes of this new viewpoint are outlined in § 4: The most important “smoking
gun” experimental probe of gravitized quantum theory is via higher order quantum inter-
ference (non-Born-rule-based) phenomena (non-Bornian quantum theory). We also discuss
how gravitational interferometry can probe the statistics of spacetime quanta. Returning
to the SM, we discuss the recently obtained seesaw formula for its hallmark mass-scale, the
Higgs mass, and in turn its intricate fermion mass and mixing structure in § 5: A cascade
of analogous seesaw formulae (see Figure 1 here and § 5.2 below) generates the entire SM-
fermion mass hierarchy (§ 5.3) and fermion mixing (CKM and PMNS) angles (§ 5.4). These
results are thus shown to impose correlating bounds on all mass-scales, the cosmological
constant (Λcc = (4πMCH /~c)2 ), the Higgs mass (MH ) as well as the masses and mixing of
quarks and leptons. Furthermore, all these results should be understood as explicit results
of quantum gravity viewed as generalized (gravitized) quantum theory. Finally, § 6 collects
our concluding comments.
3
Table 1: The iterative estimation [10] of the Standard Model mass hierarchy filigree, rewritten in
terms of the cosmological horizon scale MCH all given in (3+1-dimensional) Planck units MP . (Note
that the Yukawa couplings in the table in principle originate from the criticality of the Standard
Model as discussed in Section 5.) We thank Per Berglund for conversations on this topic.
We find that an explicit understanding of the vacuum energy (cosmological constant) problem
opens a new view on the problem of quantum gravity that also leads to novel empirical probes.
So, let us start by discussing the canonical calculations of the cosmological constant in
quantum field theory and in string theory by emphasizing various similarities and indicating
the key differences between these two calculations. We follow the presentation from our
4
MCH MP
10−34 eV 1028 eV
10−3 eV
MΛ
MSM
4.5×1023 eV
MBZ MH⋆
(0.002)
(0.06) Ue3 mν3 {m⋆t,b,τ } Vub
Figure 1: A schematic depiction of the iteratively generated mass-scales in Table 1, their horizontal
position indicating their magnitude on a logarithmic scale. Masses indicated by ⋆ involve additional
numerical parameters; see Table 1 and text. For the mixing elements, Vij and Uij separately, their
relative horizontal positions indicate their log(MCH /MP ) (· · ·) values.
Quantum Field Theory (QFT): The vacuum partition function of a free scalar in D
spacetime dimensions (which can be generalized for other fields) is
Z R 1 2 2 1
Zvac = Dφ e− 2 φ(−∂ +m )φ ∝ p , (2.1)
det(−∂ 2 + m2 )
and this expression can be rewritten as
1 2 +m2 )
Zvac = e− 2 Tr log(−∂ . (2.2)
Fourier transforming to momentum space produces −∂ 2 = k 2 , and so [19, 26]
Z
1 2 2 dτ −(k2 +m2 )τ /2
− log(k + m ) = e . (2.3)
2 2τ
Here, the Schwinger parameter, τ , is a worldline affine parameter (world-line time) of the
particle, i.e., the quantum of the field φ. Tracing then produces
Z Z
dD k 2 2 dD−1 k ωk
log(k + m ) = , (2.4)
(2π)D (2π)D−1 2
because Z Z
dτ dk 0 −(k2 +m2 )τ /2 ωk def
e = with ωk2 = k 2 + m2 , (2.5)
2τ 2π 2
where ωk is equivalent to k0 on-shell. Thereby, we arrive at the familiar result for the vacuum
energy density in D spacetime dimensions
Z
dD−1 k ωk
ρ0 = ∼ ΛD , (2.6)
(2π)D−1 2
5
identifying ΛD as the momentum space volume. This inherently divergent expression leads
to the infamous cosmological constant problem (see also [27]).3 However, as emphasized by
Polchinski, the vacuum partition function is also given as [19, 28]
with ρ0 the vacuum energy density and VD the D-dimensional spacetime volume. On the
other hand, Zvac = exp{ZS 1 }, where ZS 1 is the circle (S 1 ) partition function in the world-line
(particle) formulation:
Z Z
def dτ dD k −(k2 +m2 ) τ
ZS 1 = ZS 1 (τ ), ZS 1 (τ ) = VD e 2. (2.8)
2τ (2π)D
Combining the two expressions, the vacuum energy density becomes
iZS 1
ρ0 = ∼ ΛD . (2.9)
VD
This is an important formula that we will crucially utilize in what follows.
String Theory: Moving on to the case of a bosonic string (with obvious fermionic gener-
alizations), we can use the result for the particle vacuum energy, albeit now to an infinite
tower of particles with the stringy mass spectrum [19, 28]
2
m2 = (h + h̄ − 2). (2.10)
α′
Summing over the physical string states (“phys. st.”) then yields
X X Z Z
dr(2πr)−D/2 dθ i(h−h̄)θ − 2′ (h+h̄−2) r
ZS 1 = VD e e α 2, (2.11)
phys. st.
2r 2π
h,h̄
def
with level-matching (h = h̄, i.e., δh,h̄ ) imposed after the k-integration. Defining τ ≡ τ1 +iτ2 =
def
θ + i αr′ and q = e2πiτ as usual, the partition function of a bosonic string on T 2 (which can
be derived directly from the Polyakov path integral [19, 28]) becomes
Z X
dτ dτ̄
ZT 2 = VD (4π 2 α′ τ2 )−D/2 q h−1q̄ h̄−1 . (2.12)
2τ2 h
def
Let r = α′ τ2 , so that Z
2 ′ dD k −k2 r
−D/2
(4π α τ2 ) =e 2. (2.13)
(2π)D
Akin to the case of particles or QFT, this produces:
Z Z
def dD k 2 def dτ dτ̄
ZT 2 = VD D
f (k ) = ZT 2 (τ ) ∼ VD ΛD , (2.14a)
(2π) 2τ2
3
In D = 4 dimensions, Einstein’s equations imply the cosmological constant to be Λcc ∼ ρ0 GN ∼ ρ0 lP2 .
6
with Z Z
def dD k 2 def d2 τ −k2 α′ τ2 /2 X h−1 h̄−1
ΛD = and f (k ) = e q q̄ , (2.14b)
(2π)D F 2τ2 h
Although this is still divergent, it admits modular regularization [5,30] by restricting the sum
to a finite range5 :
4
Nq −1 Np −1 Z Z 1
X X dx dx̃ − (k+x̃) 2 ǫ2 τ
Z(τ ) = λ ǫ e 2 . (2.20)
k=0 0 2π
k̃=0
N ~4
l4 MΛ4 = N ~4 , or MΛ4 = . (2.22)
l4
2
However, e−p τ /2 ≤ 1 in (2.18) implies an upper bound: ρ0 ∼ MΛ4 ≤ N
l4
, for D = 4. Our above
2
calculation of the partition function of the bosonic string on T in D = 4 implies that the
same bound also holds in string theory:
N
ρ0 ≤ . (2.23)
l4
Stemming from the phase-space modular regularization (2.20) and as we will discuss further
in the following subsection, this result extends to quantum fields and effective potentials in
QFT, including the cosmological phase transitions (electroweak and QCD), without changing
the outcome for this bound on the vacuum energy.
4
Writing out ~ explicitly, we emphasize that ǫ and λ are, respectively, momentum- and length-scales.
5
This restriction is motivated also on physical measurement/observable grounds: spacetime integration
is limited by the finite cosmological horizon, while momenta are limited by the Planck scale since probes
of higher momentum are self-consistently invisible in the Planckian region around the event horizon of the
interacting probe-target system.
6
From the modular polarization [31] point of view, these count unit cells of vacua, not on-shell states or
particles.
8
Holography: Consider now the Bekenstein bound in a four-dimensional space with a cos-
mological horizon with the above positive cosmological constant, i.e., in de Sitter spacetime.
This spacetime metric is in static coordinates:
r2 2 dr 2
ds2dS = − 1 − 2 dt + r2
+ r 2 dωS2 2 , (2.24)
rCH 1 − r2
CH
def
where the cosmological horizon, l = rCH , is the size of the observed spacetime. Following
the discussion in [8,29], identifying the above microscopic counting of ground states with the
gravitational entropy renders the Bekenstein bound (Sgrav = lP−2 Area) as [15]7 :
l2
N≤ . (2.25)
lP2
Combining this holographic bound with the phase space bound (2.23) on ρ0 leads to
1
ρ0 ≤ , (2.26)
l2 lP2
which exhibits a mixing of the UV (lP ) and the IR (l) scales. This mixing between the short
distance and long distance scales induces a bound for the cosmological constant, which in
D = 4 dimensions (Λcc = ρ0 lP2 ) reads:
1
Λcc ≤ . (2.27)
l2
Associated with the vacuum energy density is a natural energy scale, ǫcc :
1
ρ0 = ǫ4cc ∼ . (2.28)
l2 lP2
Its corresponding natural length scale, lcc ≃ 1/ǫcc , is given by the the seesaw formula
p p
lcc ≃ l lP , i.e., MΛ ≃ MCH MP , (2.29)
where MCH ∼ 10−34 eV is the Hubble (cosmological horizon) mass-scale and MP ∼ 1019 GeV,
is the Planck scale. Finally, note that the integration over the world-line or the world-sheet
parameters does not change this final result. These integrations done naturally at the end
only provide an overall renormalization of the Newton constant.
Extending from [10], we summarize the remarkable properties of the above results:
• Using l ∼ 1027–28 m (the observed cosmological horizon) and lP ∼ 10−35 m (as a funda-
mental length unit), the seesaw relation (2.29) yields lcc ≃ 10−4 m (i.e., ǫcc ≃ 10−3 eV),
in agreement with observations and identifying ǫcc = MΛ from (2.21).
7
In our previous work on string theory in de Sitter space (section 4.6 of [9]) we have pointed out that
modular quantization and the appearance of the new quantum number N imply the holographic scaling,
and thus this IR condition is fundamentally tied to the nature of modular polarization.
9
• In the l → ∞ limit, (2.26) forces ρ0 → 0, and l is the IR length-scale. Conversely, ρ0
(and Λcc ) is nonzero and small because the cosmological horizon is finite and large.
• The relation (2.29) is radiatively stable since it is UV-independent; lP here serves as a
reference length unit.
• The extraordinary smallness of the cosmological constant thereby owes, essentially,
to the universe admitting a large number of degrees of freedom: N ∼ 10124 . The
entropic/area nature of the Bekenstein bound, (2.25), relates N to the square of the
length-ratio, l/lP ∼ 1061–62 , although N counts the number of degrees of freedom in
the 4-dimensional spacetime volume.
• In turn, the number of degrees of freedom in the universe (N) is large as that makes
scale of fluctuations small, √1N , indicating its stability.
• This estimates Ni ∼ N 1/4 ∼ 1031 (where i is t, x, y, z). Not so unreasonable in com-
parison with Avogadro’s number (1023 ) for matter degrees of freedom, it is tempting
to refer to Ni ∼ 1031 as the spacetime Avogadro number. This should be measurable
in the context of gravitational interferometry (see Section 4). From this point of view,
the characteristic scale lcc = Ni lP ≃ 10−4 m associated with the vacuum energy (and
also characteristic of certain extra dimension models) is the collective, macroscopic
spacetime scale, tied to the spacetime Avogadro number of 1031 .
Furthermore, we emphasize the quantum contextuality of the above calculation: The mea-
surement of a quantum observable depends on which commuting set of observables are within
the same measurement set of observables, i.e., quantum measurements depend on the context
of measurement. The concept of contextuality will be crucial in our analysis of masses and
mixing angles of elementary particles in Section 5.
• First, the momentum-scale ǫ is not a cut-off, since ǫ and λ can be arbitrary, albeit
reciprocally related by λ ǫ = ~; see footnote 4 on p. 8.
• Second, ǫ4 is effectively eliminated in favor of N, which is the new quantum number,
and the size of spacetime, l = rCH , the cosmological horizon, i.e., the size of the
observed classical spacetime.
• In turn, N is determined by the Bekenstein bound, (2.25), and is thereby related to l
and lP (the ultimate IR and UV scales, respectively) — which is where gravity enters,
via the familiar relation GN ∼ lP2 .
In contradistinction, EFT cannot possibly “see” N, and in particular cannot “know” about
either the Bekenstein bound or the UV/IR mixing. For example, vacuum energy routinely
cancels in the computation of EFT correlation functions. Also, EFT is defined in classical
spacetime [32,33]8 . Thus the above calculation calls for a fundamental quantum formulation
8
√
The above seesaw relation lcc ≃ l lP does appear in [34], where l and lP are, respectively, the ultimate
IR and UV length-scales in EFT, and are related by the physics of black holes/holographic bound. However,
that approach has neither of the crucial aspects of our derivation: modular representation, the number of
phase space cells N , the explicit UV/IR mixing and contextuality. We comment on the connection with
EFT at the end of Section 2.3.
10
that relies on the modular polarization [31]. Precisely this is provided by the metastring
formulation of string theory [5], which we will duly discuss in Section 3.5.
The foregoing analysis and the cosmological constant bound (2.29) extend to QFT as it com-
bines the Bekenstein bound and our phase space argument, both of which are universal and
insensitive to any QFT/EFT vacuum redefinition such as due to possible phase transitions.
Relying on standard generalizations to other fields and even string theory, consider a
scalar field theory [35, 36] with the typical Lagrangian
1 1 1
L = (∂φ)2 − V (φ) with V (φ) = m2 φ2 + gφ4 (2.30)
2 2 24
and its partition function
Z
def def
Z(J) = Dφ ei[S(φ)+Jφ] = eiW (J) . (2.31)
The generating functional of vacuum correlation functions, W (J), is a direct analogue of (2.8)
and (2.14), defines the effective action via the Legendre transform:
Z Z
def
Γ(φ) = W (J) − d x J(x)φ(x) = d4 x [ . . . − Veff (φ) + . . . ],
4
(2.32)
which in turn defines the QFT vacuum energy as the minimum of Veff (φ). (This also absorbs
the proper path integral normalization that is responsible for the vacuum energy.)
Expanding the original action S(φ) up to quadratic fluctuations and Gaussian path-
integration produces the corresponding ~-expansion of the effective action and potential,
i~
Γ(φ) = S(φ) + Tr log ∂ 2 + V ′′ (φ) + O(~2 ), (2.33)
2 Z
i~ d4 k
Veff (φ) = V (φ) − 4
log[k 2 − V ′′ (φ)] + O(~2 ), (2.34)
2 (2π)
and to the famous Coleman-Weinberg potential [37] by explicit evaluation of the momentum
integral. With the Schwinger parametrization of the logarithm [19, 26],
Z
2
dr −U (k2 ,φ) r/2
log U(k , φ) = − e , (2.35)
r
quantum corrections to the effective action may again be rewritten, most crucially, as the
phase space integral
ZZ 4 4 Z
d xd k dr −U (k2 ,φ) r/2 2
Γ(φ) ∼ 4
e with 0 < e−U (k ,φ) r/2 < 1, (2.36)
(2π) r
11
in full analogy with (2.8)–(2.14)–(2.17). Modular regularization (2.20) then again leads to
the same bound on the vacuum energy evaluated from the effective action, upon coupling
to gravity and being subject to the Bekenstein bound. Once again, the integration of the
Schwinger parameter is done at the end, and its effect is absorbed in the renormalization of
the Newton constant.
The canonical evaluation of the effective action, Γ(φ), at the effective potential mini-
mum, implies a divergent vacuum energy the cosmological constant problem upon coupling
to gravity. This evaluation is inherently sensitive to both radiative corrections and any
phase transitions dictated by the effective potential. At finite temperature T , the familiar
Landau-Ginsburg description (with a, g > 0),
1 1 1
L(φ, T ) = (∂φ)2 − V (φ, T ) with V (φ, T ) = a(T − Tc )φ2 + gφ4 , (2.37)
2 2 24
has a global minimum (φ = 0) above the critical temperature, Tc . Below Tc , L(φ, T ) develops
new global minima, to one of which the system transitions from the now unstable (tachyonic)
local maximum, φ = 0. Coupling the symmetry breaking order parameter, φ, to a gauge
field renders it massive in the classic Higgs mechanism. The above analysis however still
applies and again yields the quantum part of the effective action to scale as
ZZ 4 4 Z
d xd k dr −U (k2 ,φ,T ) r/2 2
Γ(φ, T ) ∼ 4
e with 0 < e−U (k ,φ,T ) r/2 < 1. (2.38)
(2π) r
The above-established bound for the vacuum energy (determined by the minimum of this
finite temperature effective potential) therefore continues to hold and gives the same seesaw
formula (2.29) when combined with the Bekenstein bound and modular regularization. In
particular, as standard in finite-temperature QFT, we replace
Z X Z
d4 k d3 k
(···) →T (···) , (2.39)
(2π)4 0
(2π)3
k =2πinT
which is still bounded by the volume of phase space since T measures the size of the
“imaginary time/energy” direction and the indicated summation stems from having dis-
cretized that direction in (2.39). As usual, returning to the continuum recovers the expected
P R 3 R
n d k → d4 k.
12
Throughout, the Schwinger parameter r-integration is left to the end of the calculation,
where it merely renormalizes the Newton constant, and has no influence on the vacuum
energy bound. Before the r integration, it is the phase space volume that bounds the effective
action; together with the Bekenstein bound and modular regularization, this produces the
bound on the cosmological constant as derived above. The ubiquitous appearance of the
phase space volume in these expressions for the effective action therefore justifies applying our
argument from the previous section, which together with the Bekenstein bound and modular
regularization reproduces the same seesaw formula (2.29) and cosmological constant bound,
also in the context of QFT coupled to gravity. These same characteristics of our bound imply
its radiative stability and continued validity regardless of the cosmological phase transitions
(electroweak, QCD). Conceptually (see also footnote 5), it is the mixing (2.26) of the UV
(gravitational, Planck scale) and the IR (cosmological horizon, Hubble scale) that insures
this stability. Neither local QFT nor any EFT can “see” either this UV/IR mixing or the
ensuing resolution of the vacuum energy problem, since they by construction omit any global
(non-local) features associated with modular regularization of phase space.
It is worth emphasizing that vacuum energy is in EFTs tied to the path integral nor-
malization, and so cancels in usual EFT calculations (without gravity). The standard EFT
results are recovered in a constrained double scaling limit [9], where N, l → ∞ while
def
N/l4 = 1/lΛ4 = const. < ∞, and lP → 0. This still preserves the seesaw nature of the
formula for lΛ , leading to the analogous observation of [34], where however l and lP serve,
respectively, as the EFT IR and UV cut-offs. Unlike the crucial role of quantum spacetime
degrees of freedom and the mixing between the UV and IR physics in quantum gravity, EFT
is defined in classical (and not quantum) spacetime, and so is fundamentally insensitive to
the UV/IR mixing. Nevertheless, when combined with holography, EFT can capture the es-
sential low energy features (such as the final geometric mean formula) of the fully quantum
treatment of the vacuum energy problem, that is valid both in the UV and IR regions.
In conclusion of this section, the vacuum energy (cosmological constant) problem can
be explicitly understood through a combination of phase space reasoning (with, in principle,
a dynamical phase space) and the holographic bound (as an infrared requirement). This in
turn opens a new vista on the theoretical and empirical foundations of quantum gravity.
13
used to define general quantum probabilities and observables of quantum gravity. The famil-
iar Born rule should be likened to the Minkowski metric of special theory of relativity, which
is maximally symmetric, homogeneous and isotropic, and is a linearization of the general
dynamical spacetime geometry of general relativity that has no such restrictions. Thus, the
main difference between traditional approaches to quantum gravity and the one presented in
the present review, is the gravitization of quantum theory, that is a fully dynamical geometry
and topology of the space of quantum states. In this section we present a general discussion
of first, a non-dynamical quantum spacetime (called modular spacetime) which captures the
geometry of quantum theory, and then extend this to a dynamical quantum spacetime for-
mulation found in a non-commutative and phase-space like formulation of string theory (the
metastring).
In this subsection we present a model of quantum spacetime, called modular spacetime based
on the most general geometry of quantum theory, dubbed Born geometry. Born geometry
(the geometry of quantum relativity, and a unification of symplectic, doubly orthogonal and
doubly metric geometry) should be understood as a direct analog of Minkowski geometry
in the context of classical relativity. As a matter of fact, we suggest the following analogy
between classical and quantum relativity.
Classical relativity can be understood as a logical progression from a) special relativity
— motivated by classical field theory — (with Minkowski spacetime/geometry and relativity
of simultaneity) to b) relativistic field theory (with, in the quantum context, unitary repre-
sentations of the Lorentz/Poincaré symmetry and with the famous prediction of antiparticles
and the spin-statistics relation) and, finally, c) general relativity (with dynamical classical
spacetime). Here, spacetime relativity is the first (classical) relativity (with both spacetime
and matter being classical). General relativity is a dynamical extension of the same.
Quantum relativity can be analogously understood as a logical progression from A) Quan-
tum mechanics (QM) understood from quantum spacetime (with modular spacetime, Born
geometry and relative (observer dependent) locality to be explained in what follows) — this
is a Bornian quantum theory; to B ) QFT (with metafields and the new prediction — meta-
particles — this is an intrinsically non-commutative covariant formulation based on modular
spacetime with both spacetime and dual spacetime as natural limits) — this is also a Bor-
nian quantum theory; and, finally, C ) gravitized quantum theory — a quantum analog
of general relativity, a non-Bornian quantum theory (with dynamical quantum spacetime
and dynamical Born geometry as realized in the metastring formulation of an intrinsically
non-commutative, T-duality covariant and (chiral) phase-space-like string theory with meta-
particle zero modes. Metaparticles appear as natural dark matter quanta and the geometry
of dual spacetime as the natural origin of dark energy. Similarly, general quantum statis-
tics of spacetime quanta, as well as general higher order quantum correlations characterize
14
gravitized quantum theory, which as a sort of metaquantum theory, also sheds light on the
origin of QM and QFT.) Here Quantum Mechanics (QM)/Quantum Field theory (QFT)
can be viewed as second (quantum) relativity (with matter being quantum, and spacetime
classical). In the same vein, Quantum Gravity (QG) = Gravitized Quantum theory (GQ)
can be viewed as third (quantum gravity or gravitized quantum) relativity (with both space-
time and matter being quantum). (“Third relativity” is apparently the phrase advocated by
Finkelstein and Wheeler; see David Finkelstein’s book [38].)
The space of commuting subalgebras of the Heisenberg algebra, [q̂, p̂] = i~, which in the co-
variant (self-dual lattice) phase space formulation becomes the modular spacetime [5, 30, 31]
is the target space of the metastring [5] and the metaparticles [40–42] to be explained in
the next subsection. For example, vertex operators in metastring theory are representations
of this Heisenberg algebra. This description (intrinsically non-commutative, since [x, x̃] = i,
where x ≡ q/λ and x̃ ≡ p/ǫ and λǫ = ~) will appear in what follows in the metastring formu-
lation of string theory which avoids all of the co-cycles that turn up in standard descriptions
of the vertex operator algebra in string theory [19, 20], the vertex operators being the above
generators of translations in phase space.
A more elementary (and familiar) argument for the existence of modular spacetime may
be presented as follows: In quantum theory, short (UV) distances are associated with high
energy, as implied by the indeterminacy relation, δq ∼ 1/δp (in ~ = 1 units). On the other
hand, in classical (as well as semiclassical) gravity, the Schwarzschild radius RS of a mass M
is given by RS ∼ GM ∼ lP2 M, where G ∼ lP2 is the gravitational constant in 4-dimensional
spacetime, with lP the Planck length. In quantum gravity, quite generally, one therefore
expects that higher energy leads to larger (IR) distances δq ∼ lP2 δp. These diametrically
contrasting behaviors (associated with UV and IR) may be reconciled by relating the UV
and IR physics: Recall that, given a fundamental lattice length, quantum states are de-
scribed in terms of quantum numbers associated with both a lattice and its dual [31]. In
our present case, this involves momenta p and their duals p̃, provided that these commute
[p, p̃] = 0. The indeterminacies δp and δ p̃ thereby being interchangeable provides the first
substitution in the chain: lP2 δp ∼ δq → lP2 δ p̃ ∼ δq → lP2 (δ q̃)−1 ∼ δq ⇒ δq δ q̃ ∼ lP2 , where
15
the second replacement used the canonical δ p̃ δ q̃ ∼ 1 indeterminacy relation. This implies
a new fundamental non-commutativity between spacetime and dual spacetime coordinates
[q, q̃] ∼ ilP2 . The commutative nature of modular variables in quantum theory insures that
this can be completely covariantized [31]. Thus, combining the fundamental quantum and
gravitational relations between spatial distances and momenta leads to:
Note that modular variables are covariant (there is modular energy and modular time as
well). Take the fundamental length λ and energy ǫ, so that λǫ ≡ ~. Modular variables are
non-local (but consistent with causality — this is the origin of the uncertainty principle). We
emphasize the notion of contextuality: in a double slit experiment the parameters λ and ǫ are
contextual to the experiment. This contextuality was an important point in our discussion
of the cosmological constant problem and it will be important in Section 5 when we discuss
masses and mixing angles of quarks and leptons. Also we stress the explicit non-locality:
p2
Take H = 2m + V (q) and write the Heisenberg equation of motion for eipR/~ , or equivalently
[p]R . (Here R is a contextuality parameter, such as the distance between two slits in the
double slit interference experiment.)
Now, let us reformulate quantum mechanics (QM) using (covariant) modular variables via
modular spacetime. (Quantum theory tells us something new about quantum spacetime via
the concept of modular spacetime.) What precisely is modular space? Modular space is the
space of all commuting subalgebras of the Heisenberg-Weyl algebra. By definition [q, p] = i~
is the Heiseinberg-Weyl algebra, whereas [[q]a , [p]2π~/a ] = 0 is the commuting subalgebra of
Weyl-Heisenberg. Here we have the following fundamental result encapsulated in Mackey’s
Theorem: the space of all commuting subalgebras of the Heisenberg-Weyl algebra is a self-dual
phase space lattice lifted to the Heisenberg-Weyl algebra. This theorem allows us to define
modular spacetime and the associated Born geometry.
In particular, if we use covariant modular variables we obtain modular spacetime of d
spacetime dimensions. In the above theorem the concept of phase space comes with the
16
natural symplectic structure Sp(2d), ωab . The concept of a self-dual lattice (ℓ ⊕ ℓ̃), where ℓ
is a lattice implies doubly-orthogonal O(d, d), ηab . Finally, to define the vacuum on this self-
dual lattice, we need doubly metric structure O(2, 2d − 2), Hab . The triple (ω, η, H) defines
Born geometry. Their triple intersection gives the Lorentz group. Thus QM follows from
non-locality (fundamental length/time of quantum gravity) that is consistent with causality
(implied by the Lorentz symmetry). We emphasize that one can be localized (and thus local
QFT is possible) in a particular phase space cell, but one can not tell in which phase space
cell (this is the origin of the uncertainty principle), because the number operators (for the
spatial and the momentum directions, respectively) do not commute with modular variables.
How can fundamental length/time be consistent with the Lorentz symmetry? (In some
sense, this is one of the main puzzles of quantum gravity.) This is possible because of
relative (observer dependent) locality [43]). Different observers see different spacetimes (slices
of modular spacetime). However, different spacetimes are in linear superposition, and so
fundamental length/time is consistent with the Lorentz symmetry. (This is similar to spin:
the superposition of up and down spin gives the Bloch sphere which is consistent with
rotation symmetry, even though spin is discrete [31].) Thus linear quantum superpositions
are needed to reconcile the fundamental length/time with the Lorentz symmetry. This is
the quantum spacetime origin of quantum theory.
Next we introduce the generic quantum polarization — modular polarization (defined
via the Zak transform). Given Schrödinger’s ψn (x) define the modular wave function
√ X −2πinx̃
ψλ (x, x̃) ≡ λ e ψn (λ(n + x)), (3.4)
n
(x ≡ q/λ, x̃ ≡ p/ǫ, so [x, x̃] = i, λǫ = ~). From the point of view of modular polarization,
Schrödinger’s polarization is very singular. Introduce XA ≡ (xa , x̃a )T , so that [X̂a , X̂b ] =
iω AB . We can write the translations operators in phase space covariantly WK ≡ e2πiω(K,X) ,
where K stands for the pair (k̃, k) and ω(K, K ′) = k·k̃ ′ −k̃·k ′ . (W should be really understood
as Aharonov-Bohm phases, which are prototypical examples of modular variables.)
So far we have discussed covariant quantum phase space as an examples of modular
space, and so we are ready to discuss modular spacetime. Consider [42] a metaparticle (mp)
propagating in a modular space defined by Born geometry, (ω, η, H). The metaparticle
R
world-line action Smp = dτ Lmp (with the canonical particle emerging in the µ → 0 and
p̃ → 0 limit), with Lmp = pµ ẋµ + p̃µ x̃˙ µ + λ2 pµ p̃˙ µ − N2 (pµ pµ + p̃µ p̃µ − m2 ) + Ñ (pµ p̃µ − µ) ,
where ω is in (“the Berry-phase”) pµ p̃˙ µ , and η in the diffeomorphism constraint pµ p̃µ = µ
and H in the Hamiltonian constraint pµ pµ + p̃µ p̃µ = m2 . Here it is natural to talk of dual
spacetime x̃, [x, x̃] = iλ2 , and dual momentum space p̃, [p, p̃] = 0. (Also, [x, p] = i~ = [x̃, p̃].)
The metaparticle can be understood also as follows: If one second quantizes Schrödinger’s
ψ(x) one naturally ends up with a quantum field operator φ̂(x). Similarly, the second quan-
tization of the modular ψλ (x, x̃) would lead to a modular quantum field operator φ̂λ (x, x̃)
(modular fields or metafields)
φ̂(x) → φ̂λ (x, x̃), (3.5)
17
with [x, x̃] = iλ2 defining a covariant non-commutative field theory [40, 41] that depends on
the contextuality parameter λ. Classical spacetime label x of canonical QFT corresponds
to a choice of (classical spacetime) polarization in modular (quantum) spacetime with a
contextuality parameter λ. Quanta of canonical quantum fields φ(x) are particles (and
their antiparticles). Similarly, quanta of modular quantum fields φλ (x, x̃) are metaparticles.
Thus, the first prediction of modular spacetime approach to quantum theory concerns the
existence of metaparticles [42]. (We will argue that dual particles, correlated to visible
particles, represent dark matter.) If we turn on backgrounds p → p + φ and p̃ → p̃ + φ̃. Thus
we have “dark matter” fields, φ̃(x), in the effective classical spacetime x description (after
integrating over the dual spacetime x̃). However, the visible φ and invisible (dark matter) φ̃
do not commute. Therefore, one could say that dark matter is fuzzy from the point of view
of classical spacetime.
A few other comments are in order [5, 31, 44]: modular spacetime has double the dimension
of spacetime. The modular cells are not simply connected (there is a unit flux through each
cell — this could be considered as the origin of matter/fermionic degrees of freedom; the
deformations of each cell could be understood as the origin of bosonic degrees of freedom,
i.e. interaction quanta). We note the possibility of general quantum statistics and general
higher order quantum-correlations, to be discussed in Section 4. Similarly, one needs to
adopt double scale (UV/IR) Renormalization Group characteristic of non-commutative field
theory [44] in order to define the continuum limit as well as renormalization of couplings and
correlation functions in the context of modular quantum field theory. In some sense, modular
quantum field (metafield) theory is the limit of the underlying quantum gravitational origin
of quantum field theory defined in quantum spacetime, with the local (Wilsonian) quantum
field theory in a fixed spacetime background, being a singular, classical spacetime limit, of
this more fundamental description.
We emphasize that modular wavefunctions are quasiperiodic [31]. Classical spacetime
emerges from the process of extensification (imagine one unit length in dual direction and
many, N, modular cells in the spacetime direction). Spacetime emerges, in the large N limit,
as a natural pointer basis in quantum theory [31]. (And this sheds new light on the problem
of quantum measurement. The dual spacetime labels can be understood as covariant “hidden
variables” associated with the spacetime pointer basis, selected by quantum gravity.) Also,
spacetime and matter appear as ‘two sides of the same coin”. Similarly, cosmology is an
interplay between visible and dual (invisible) degrees of freedom [45].
One can compute the propagator for the metaparticle [42]
δ(p · p̃ − µ)
G(p, p̃; pi , p̃i ) ∼ δ (d) (p − pi )δ (d) (p̃ − p̃i ) . (3.6)
p2 + p̃2 + m2 − iε
18
The canonical particle propagator is a highly singular p̃ → 0 (and µ → 0) limit of this
expression. One also obtains the following dispersion relation (in a particular gauge p̃~ = 0)
µ2
Ep2 + = ~p2 + m2 . (3.7)
Ep2
For each particle at energy E there exists a dual particle at energy Eµ . (Analogous to
the prediction of antiparticles in QFT.) This dispersion relation is indicative of quantum
gravity phenomenology in the infrared limit, thus contradicting the usual intuition about
the relevance of the Planck scale for quantum gravity. (Given the relevance of quantum field
theory in the context of many body physics, one might wonder if metafields and metaparticles
find their use in that domain of physics as well. Indeed, one can consider non-relativistic
metafields and their metaparticle quanta and one can, for example, derive a Friedel-like
static potential for metaparticles and also introduce the concept of quasi-metaparticles in
the realm of many body physics [46].)
Note that dual “particles” (dual fields) are natural candidates for dark matter because
to leading order in λ
Z p
Seff = − g(x)g̃(x̃)[R(x) + R̃(x̃) + Lm (A(x, x̃)) + L̃dm (Ã(x, x̃))]. (3.8)
Here the A fields denote the usual Standard Model fields, and the à are their duals, as
predicted by the general (modular) formulation of quantum theory that is sensitive to the
minimal length. In the above expression one needs to integrate over the dual space coor-
dinates x̃ to get an effective description of visible matter, A(x), and dark matter, Ã(x), in
classical x spacetime.
Similarly, dynamical geometry of dual spacetime represents the natural origin of dark
energy (again, to leading order in λ)
Z p p
Seff = − −g(x) −g̃(x̃)[R(x) + R̃(x̃) + . . . ]. (3.9)
In this leading limit, the x̃-integration in the first term defines the gravitational constant
GN , and in the second term produces a positive cosmological constant constant! In general,
visible and dark matter degrees of freedom are correlated (via the minimal length λ). This
suggest the origin (from dark matter) of the observed scaling (found in galaxies, clusters,
superclusters) and the universal acceleration a0 ∼ cH/(2π) (with the observed positive
cosmological constant Λ ∼ H 2 ). Metaparticles can be understood as fuzzy dark matter,
which in turn does point to a natural relation of vacuum energy (dark energy) and fuzzy
dark matter [47, 48].
def
Here, Xa = (X a /ℓs , X̃a /ℓs )T are coordinates on phase-space like (doubled) target spacetime
and the fields η, H, ω are all dynamical (i.e., generally X-dependent) target spacetime fields9 .
In terms of the left- and right-moving 0-modes of the twenty-six dimensional bosonic string,
one defines
def def
xa = xaL + xaR , x̃a = x̃aL − x̃aR . (3.11)
In the context of a flat metastring, the coefficients ηab , Hab and ωab 10 are constant:
0 δ h 0 0 δ
ηab = , Hab = , ωab = , (3.12)
δT 0 0 h−1 −δ T 0
where h denotes the flat (1, d−1)-dimensional metric and δ is the Kronecker symbol. The
standard Polyakov action is then obtained by setting ωab = 0 and integrating out the x̃a ,
Z
SP = dτ dσ γ αβ ∂α X a ∂β X b hab + . . . (3.13)
The triplet (ω, η, H) defines Born geometry [5, 13] (which is ultimately dynamic, suggesting
a “gravitization of quantum theory” [53–55]) so that the metastring propagates in a (dynam-
ical) modular spacetime, a phase space like structure that naturally arises in any quantum
theory [31], as argued in the previous subsection. One of the key consequences of this is that
the metastring is intrinsically non-commutative and also that its low energy QFT-like de-
scription in modular spacetime is intrinsically non-commutative. Thus every Standard Model
field φ(x) is doubled as φ(x, x̃) and φ̃(x, x̃), with doubled and non-commutative arguments
[xa , x̃b ] = iℓ2s δ a b . The quanta of such modular fields are the zero modes of the metastring —
the metaparticles — whose dynamics, as already discussed, are given by a world-line action
involving a doubling of the usual phase space coordinates. The metaparticle (“mp”) action
9
Recall that the ten-dimensional superstring comes out of the twenty-six dimensional bosonic string [50].
Similarly, the matrix formulation of M-theory in eleven dimensions would emerge from a non-perturbative
matrix formulation of the metastring [9] (notice the lack of the overall trace — this is a matrix model as matrix
quantum mechanics
of Born-Heisenberg-Jordan), and it should be understood as gravitized quantum theory,
R
Sstr = dτ ∂τ X̂ gACD (X̂) − [X̂A , X̂B ]hABCD (X̂) [X̂C , X̂D ] , where A, B, C, D run from 0, 1, . . . 26. Note
non A
that the first term is Chern-Simons-like and the second Yang-Mills-like, both being of the open string origin.
The fully compactified non-perturbative bosonic metastring provides the basic elementary constituents for
extensifications [5,31] from zero dimensions to four quantum spacetime dimensions (natural from the point of
view of string cosmology [51]). Finally, we note an interesting structural connection of the non-perturbative
metastring to the “palatial” (non-commutative) twistor theory [52]. The intrinsic time asymmetry of the
non-perturbative metastring formulation might be the origin of an intrinsic gravitational CP violation.
10 ch
Setting ωab = 0 directly connects Sstr to double field theory.
20
describes the zero modes of the metastring and is of the form which is fixed by the three
ingredients of the Born geometry [40–42, 45]
Z 1 h i
def N 2
Smp = ˙ ′ ˙
dτ p · ẋ + p̃ · x̃ + α p · p̃ − 2 2
p + p̃ + m + Ñ (p · p̃ − µ) , (3.14)
0 2
where the dot-product denotes contraction with signature (−, +, . . . , +). The new feature
here is the presence of a non-trivial symplectic form on the metaparticle phase space, the
non-zero Poisson brackets being
{pµ , xν } = δµν , {p̃µ , x̃ν } = δµν , {x̃µ , xν } = πα′ δµν , α′ ∼ ℓ2s . (3.15)
21
have co-cycles in the Polyakov string if we assume that [x, x̃] = 0 [19]
The metastring has dynamical Born geometry, ωab (X), ηab (X), Hab (X), but Born geometry
is the geometry of the modular spacetime formulation of quantum theory. Thus by making
Born geometry dynamical we can “gravitize quantum theory” (that is, make the geometry
of quantum theory dynamical) [13, 53]. The metastring is a theory of quantum gravity, and
so we arrive at the advertised dictionary “quantum gravity = gravitized quantum theory”.
In what follows, we argue that triple and higher order quantum interference is one of the
central observational consequences of this dictionary. This reasoning is “top-down”.
Recall that classical classical gravity (general relativity) gravitizes all of classical physics,
but making the relevant fundamental equations generally covariant, and thus it is not strange
to expect that quantum gravity requires gravitization of quantum theory. Therefore, con-
sider particle interactions as 0+1 quantum gravity, by remembering that quantum field
theory can be understood as 0+1 quantum gravity/cosmology. (This discussion could be
generalized for the case of string field theory and its fundamental cubic vertex, the “pants
12
Work in progress by Nikolina Ilic, Dejan Stojkovic, Doug Gingrich, Luca Colangeli, Mathias Roman,
Sebastien Roy-Garand, Yun Qing Wu and DM.
13
We thank Per Berglund, Andy Geraci and Dave Mattingly for conversations on this topic.
22
diagram”, by viewing string theory, including its metastring formulation, as 1+1 quantum
gravity/cosmology.) For example: consider the φ3 theory. The relevant classical equations
read
(∂ 2 + m2 )φ + gφ2 = 0 (4.1)
Here we should understand the classical field φ as the wave-function of 0+1 universes.
Thus the above classical field equation should be understood as a non-linear Wheeler-
DeWitt (WdW) equation. The interaction vertex is indicative of topology change. We have
the classical spacetime viewpoint where the decay of a particle (or scattering of particles,
the S-matrix) is viewed by utilizing the Born rule. However, the QG=GQ viewpoint from
0+1 universes (particles) is that there exists an intrinsic triple correlation. This is another
motivation for “gravitization” of quantum theory. For example, the fundamental triple vertex
(the pants diagram) of string theory becomes the fundamental triple quantum correlation
of 1 + 1 dimensional quantum gravity. (Also, such gravitization of quantum theory, in the
context of the metastring formulation of string theory, sheds new light on the background
independent definition of string field theory.)
In general, if we view general relativity as a theory of interacting gravitons (closed
strings), we have n-correlations; n = 3, 4, 5 . . . and thus the fundamental object of general
relativity, the classical spacetime, hides higher order (and in principle infinite order) quantum
correlations. By reinterpreting the non-linear interactions terms in the non-linear WdW
equation as the time evolution operator, time could be understood as measuring the rank of
general quantum correlations, and space as the size of those correlations. This sheds a new
light on the problem of time in quantum gravity, as well as at the fundamental question of
the expanding spacetime in the context of cosmology. The fundamental quantum gravity
description contains all quantum correlations, and the emergent classical spacetime hides all
correlations except for the maximally symmetric (Born-like) correlations associated with the
matter degrees of freedom. Thus quantum matter that exists in classical spacetime follows
the Born rule and canonical quantum theory. The classical world, or the results of quantum
measurement, represents the memory of those higher order quantum correlations (topology
change in the space of quantum states) hidden in the classical spacetime “condensate” in
which the canonical quantum theory operates, with definite results of quantum measurements
in such classical spacetime.
Apart from the top-down rationale, there exists a “bottom-up” reason for “gravitization
of quantum theory” [63–66]. The canonical geometry of quantum theory (as reviewed by
Ashtekar and Schilling [67] as well as [68]) is encoded in a maximally symmetric geometry
of complex projective spaces (defined by a symplectic structure, compatible with the metric
structure — the Born rule — and the product of the symplectic and metric structure that
defines the complex structure, ultimately responsible for quantum interference). The quan-
tum clock relates the Born rule (the Fubini-Study metric of complex projective spaces) to
infinitesimal time [69] (see also Aharonov’s earlier work [70]),
P2 (A, B)+P2 (B, C)+P2 (C, A)−P1 (A)−P1 (B)−P1 (C), (4.6)
where only pairwise interferences between the pairs (A, B), (B, C), and (C, A) appear. It is
clear from the above that in order for there to be a non-linear correction in an interference
pattern the Born rule must be relaxed. Consider a triple slit experiment: Since only pairwise
interferences between the pairs (A, B), (B, C), and (C, A) appear, it makes sense to define
any deviation from this relation as the intrinsic triple-path interference I3 (A, B, C)
(This can be easily generalized for the case of n-paths.) For both classical and quantum
theory, this intrinsic triple-path interference is zero for any triplet of paths. Experimental
confirmation of I3 = 0 would be a confirmation of the Born rule. Weak bounds were placed
on the parameter (κ ∼ 10−3 ) in photonic experiments (see the references in [72])
ε
κ= , ε = I3 (A, B, C), δ = |I2 (A, B)| + |I2 (B, C)| + |I2 (C, A)|. (4.8)
δ
The claim of [72] is that with quantum gravitational degrees of freedom turned on, one
can get I3 6= 0, but for that one needs gravitized quantum theory, with observer dependent
Hilbert spaces and dynamical Born rule. Inspired by metastring theory, the generalized
probability in this approach to quantum gravity is given by
where a, b, c are state-space indices and with (schematically). One non-relativistic quan-
tum gravity model is provided by the canonical Schrödinger dynamics perturbed by Nambu
quantum theory [72] (in the non-relativistic limit)
dψa
= Γabc ψb ψc , (4.10)
dτ
where τ is the appropriate evolution parameter (and higher order generalizations dψ dτ
a
=
Γabcd ψb ψc ψd , etc. Here Γabc is such that one has Schrödinger’s evolution for a fixed Hilbert
space.) Note that the Schrödinger equation can be understood as a geodesic equation on
complex projective spaces, which are Einstein’s spaces (maximally symmetric, homogeneous
and isotropic). The above generalized evolution can be understood to originate from the self-
dual nature of the equations of motion of the metastring [13] (a fully relativistic gravitized
25
quantum theory). Thus even the generalized quantum evolution equation can be understood
as a geodesic equation of more general Einstein-like equations on the spaces of quantum
states with general geometry and topology. These in turn are consistent backgrounds of the
non-perturbative metastring endowed with fully dynamical Born geometry.
Here we comment that the above model of Nambu quantum theory [66,75] is essentially
based on volume preserving diffeomorphisms where the generator of volume preserving trans-
formations is caused by the Nambu bracket [76], a generalization of the Poisson bracket. The
classic example is the asymmetric top which can be re-written as a model of Nambu’s clas-
sical mechanics. The Nambu quantum theory can be understood in the above Schrödinger
representation or in the matrix representation which requires cubic and higher order ma-
trices [77]. In gravitized quantum theory complex projective spaces are generalized (for
example: the Bloch sphere becomes a Riemann surface of infinite genus) where higher order
quantum correlations are indicative of a dynamical Born rule (with handles representing
higher components ψ3 , where ψ1 and ψ2 are the real and imaginary parts of the canoni-
cal complex wave function). The classical limit is captured by topological branching (the
quantum metric, or, equivalently, probability, becomes degenerate and equal to zero). A
re-summation of the infinite number of multilinear extensions results in a general-relativity-
like theory in the general space of states (that includes ψ3 etc). The canonical quantum
theory is the maximally symmetric limit of this more general and “gravitized” formulation
of quantum theory (emerging by averaging over the infinite number of handles).
Notice that effective triple interference is possible in non-linear optical media [78] and
this experiment provides another motivation for our work [72]. (In that context, instead of ψ
we have non-linear waves and instead of probability P — non-linear/cubic energy density.)
The “smoking gun” experiment for gravitized quantum theory is thus the Talbot effect on
a diffraction grating that is turned into a non-linear Talbot effect [72]. This intrinsic triple
interference with quantum gravity degrees of freedom is analogous to the model of quantum
spacetime as a non-linear “quantum spacetime medium”. (Here we stress that no non-linear
quantum theory with fixed Hilbert spaces can possibly contend to be a gravitized quantum
theory.) Based on the discussion of the vacuum energy/cosmological constant (possibly
the first experimentally detected effect of quantum gravity) it can be argued that quantum
gravity effects appear at low energy scales: such as the cosmological constant scale 10−4 m
or or the natural particle physics scale 10−19 m.14
26
observation about the atomic structure of matter. Matter is granular and cuttable: it consists
of fermions that are held together through interactions that are mediated by bosons (the
spin-statistics theorem of local QFT). In order to extend this atomic picture to spacetime
we note the fundamental difference between spacetime and matter: spacetime is extended
and non-cuttable. In what follows we claim that spacetime quanta obey infinite statistics
and are held together by higher order quantum correlations responsible for higher order
interference effects (already discussed in the previous subsection). Thus, classical spacetime
(with all its features) is a left-over from the quantum gravity phase, which involves all higher
order (triple and higher) quantum correlations. On the other hand, matter is captured by
degrees of freedom with only Born-like quadratic and maximally symmetric correlations of
canonical quantum theory in a background of such an emergent classical spacetime. Classical
spacetime also provides the pointer basis for quantum measurements involving matter degrees
of freedom, with an inevitably classical nature of such measurement outcomes.
We argue15 that the quanta of quantum spacetime defined as modular spacetime obey
infinite statistics. Using infinite (or quantum Boltzmann) statistics we derive the fluctuation
of (modular) energy discussed by Verlinde and Zurek [82] and by Zurek in [83]. In what
follows we will use the basic formulae of the thermodynamics of infinite statistics from
Section 3 of Ref. [84], and references therein (see also, [85, 86]).
Given the fundamental commutator, [x, x̃] = ilP2 , between spacetime x and dual space-
time x̃, discussed in Section 3, both x and x̃ have to be infinite matrices. Thus their
probability distribution has be governed by non-commutative probability theory or equiva-
lently by quantum distinguishable, or quantum Boltzmann statistics, also known as infinite
statistics. This statistics (which, in the simplest case, is the analog of the Gaussian statistics
for non-commutative variables) and which is studied as such in non-commutative proba-
bility theory of Voiculescu and collaborators [87] and reviewed in the paper by Gross and
Gopakumar [88] can be applied to modular cells of modular spacetime, which are quantum
and distinguishable.
Note that Strominger [89] pointed out that the concept of infinite statistics is relevant for
black holes, and this fits naturally with the claim that the statistics of “spacetime atoms”, as
represented by modular cells, is quantum statistics of distinguishable objects. Modular cells
are quantum and they are distinguishable by construction, and thus it is not surprising that
they should obey quantum distinguishable or quantum Boltzmann statistics, that is, infinite
statistics, the natural covariant statistics of extended objects. In fact, Greenberg [90, 91]
has pointed out that infinite statistics is the only statistics consistent with non-locality
and CPT/Lorentz symmetry. This again is consistent with the basic features of modular
spacetime cells — modular spacetime is covariant, but modular cells are non-local objects.
(Also, modular spacetime is the background of the metastring formulation of string theory,
a covariant theory of non-local objects — strings. In principle, it should be possible to prove
a stringy version of spin-statistics theorem in string field theory, which would naturally lead
15
We thank Laurent Freidel, Jerzy Kowalski-Glikman and Rob Leigh for conversations about this topic.
27
to infinite statistics of strings.)
In what follows we present a summary of basic concepts from [84] (and the references
therein). Infinite statistics is defined in terms of a free algebra (Cuntz algebra) of operators
ai |0i = 0. (4.12)
Note that, unlike the Bose-Einstein (BE) or Fermi-Dirac (FD) statistics for matter (bosonic
or fermionic) degrees of freedom, we do not have a commutator or anticommutator relation
between a and a† . The defining equation (4.11) is the exceptional q = 0 case of the q-
deformed (quon) commutator
ai a†j − qa†j ai = δij , (4.13)
where BE corresponds to q = 1 and FD to q = −1, in both of which quanta are indistin-
guishable. In contradistinction, the exceptional q = 0 case specifies the algebra (4.11) for a
quantum Boltzmann statistics, that is, a distinguishable quantum statistics.
The thermodynamics of a system of particles obeying distinguishable quantum statistics
was studied in [92] (sections 3 and 4). These calculations were motivated by the discussion
of Schwarzschild black holes as bound states of D0-branes in Matrix theory, a light-cone,
holographic, formulation of M-theory [93], as explored in [94]. The partition function Z =
P −βEi
ie of a free gas (in d − 2 transverse dimensions of spacetime) of N particles obeying
infinite statistics with the leading order free Lagrangian L = NR
v 2 is [92]
where V is the volume and T ∼ β1 temperature, and R is a characteristic length scale that
sets the relevant mass/energy scale. (Corrections to the free Lagrangian of the form v 4 /r n
and v 6 /r 2n , where v denotes the velocity and r distance, can be also taken into account in
this calculation [92].) Notice how this differs from the usual expression for the classical par-
tition function which includes the factor (1/N!). This is a distinguishing feature of quantum
Boltzmann statistics (as pointed out in [90, 91]).
Given the above expression for the partition function all other thermodynamic functions,
including energy and entropy, are determined. How can this expression be reconciled with
non-extensive nature of entropy (holographic scaling) in the context of gravity? In order
to obtain the holographic scaling for entropy we need a relation between R and the size of
the gravitational system, for example, a black hole. The partition function of a free gas
of distinguishable particles in a volume V ∼ bd−2 can be matched to the thermodynamic
properties of a Schwarzschild black hole in d spacetime dimensions with the Schwarzschild
radius b [92]. (Note that d − 2 denotes the number of transverse dimensions, as required by
the holographic scaling.) As pointed out in [92], in the limit in which the thermal wavelength
28
√ 1
λ = βR of such a gas is of the order of the Schwarzschild radius λ ∼ b ∼ V d−2 , the partition
function Z, which can be rewritten as
V N
Z∼ , (4.15)
λd−2
and therefore, generically, log Z ∼ N, because λd−2 ∼ V . (Note that this form of Z follows
from a direct computation of the canonical partition function for the case of infinite statistics,
without any reliance on holography and black hole physics, as discussed [95], and references
therein.) In that limit, the temperature T ∼ β1 and the size b match the formulae for the
temperature and the Schwarzschild radius, respectively, from black hole thermodynamics
[92]. In particular, the energy E is calculated as [92]
∂ log Z N
E=− ∼ , (4.16)
∂β β
and the entropy S is given as [92]
S = log Z + βE ∼ 2N ∼ N. (4.17)
where the numerical factor in front of N is not important for our central point for very large
N (as is the case in our context). Thus, we obtain that the entropy is proportional to the
number of infinite statistics particles [92]
S
S∼N →E∼ . (4.18)
β
1
As already pointed out, the requirement log Z ∼ N amounts to the condition λ ∼ b ∼ V d−2
[92], or more explicitly
V (T /R)(d−2)/2 ∼ 1, (4.19)
which, after using the interpretation of the above energy E as the light-cone energy related
R
to the mass M of a boosted object E = N M 2 [94] (realized in Matrix theory, a light-cone
formulation of M-theory) implies that the entropy scales as [92]
S ∼ (T /R)−(d−2)/2 , (4.20)
and this is indeed true for Schwarzschild black holes and thus it embodies the holographic
principle (in other words, the fact that the black hole entropy scales with the area of the
horizon). We also note that this argument can be extended to the cosmological horizon, and
the holographic scaling in that case, which was relevant in our recent computation of the
bound on the vacuum energy [8].
The fluctuation of energy (the central quantity computed by Einstein in his seminal
papers on early quantum theory and statistical physics) is another derivative with respect
to β. To be more precise given
1X
hEi = Ei e−βEi , (4.21)
Z i
29
then the fluctuation
hǫ2 i ≡ hE 2 i − (hEi)2 , (4.22)
is determined as follows
∂hEi
hǫ2 i = − . (4.23)
∂β
From the expression for the energy we obtain that
N E
hǫ2 i ∼ 2
∼ , (4.24)
β β
which then implies the result [82] about the Brownian-motion-like geometric mean
s
p E
hǫ2 i ∼ . (4.25)
β
The natural temperature T ∼ β1 for a gas of spacetime atoms obeying infinite statistics, is
by construction, of the order of the Planck temperature (or equivalently, the Planck energy).
Thus, if we invoke the IR properties of quantum gravity, according to which higher energies
E and momenta correspond to larger distances l, then we can write E ∼ l, so that the
above equation becomes δ 2 ∼ l lP , (another crucial equation from [82]) where δ denotes the
fluctuation of length, associated with the relevant fluctuation of energy. We will comment
on the phenomenological implications of this equation in the next subsection.
Note that this relation is true in any number of dimensions, whereas, as discussed in the
following subsection, our recent formula (reviewed in Section 2) for the vacuum energy scale
is dimension dependent. (However, in 4 spacetime dimensions the two expressions coincide.)
Also, this result is based solely on statistics and not on any particular model of quantum
gravity. At this point one might ask if given the above result one can go back and deduce
that infinite statistics is inevitable. Given what is known about the quon statistics [95],
this statement is true: the only distinguishable statistics that supports the above relation
between the fluctuation of energy and its average is indeed infinite statistics. Also, given the
available knowledge of the thermodynamics of infinite statistics [95], it is clear that infinite
statistics cannot be associated with any “material” (i.e., matter degrees of freedom), but
that does not prevent it from being associated with the statistical properties of spacetime
atoms.
This result is interesting because of the fact that modular cells are quantum objects
and they are distinguishable and thus their statistics is quantum Boltzmann, or infinite
statistics. Thus if one models quantum spacetime as a free gas of objects obeying infinite
statistics (quanta of spacetime, with covariance being implied by the quantum nature of
this description, as discussed in § 3) then one does obtain the result of [82] in any number
of dimensions for this particular model of quantum spacetime. (We emphasize a pleasing
picture: the quantum matter degrees of freedom are fermionic, the quantum degrees of
freedom of interaction forces are bosonic, and the quanta of spacetime obey infinite statistics.)
30
From this point of view one could propose the use of gravitational interferometry to
probe this particular modular quantum “spacetime foam”, and the above unique features
of infinite statistics of modular cells (spacetime atoms). This would be a clear analogy of
Einstein’s (and Smoluchowski’s and Langevin’s) theory of Brownian motion and its exper-
imental probe by Perrin [96–99], however, in the domain of quantum gravity. A concrete
proposal is as follows: Start with a theory formulated in modular spacetime, such as the
metastring. Then we have natural matrix-like spacetime variables (say for the metastring
with a modular world sheet). Consider a de Sitter spacetime (with a characteristic scale —
the Hubble scale) as a bound state in such a theory (this is possible because the metastring
theory is bosonic, and the metastring supports the computation of the vacuum energy re-
viewed in Section 2. de Sitter spacetime has the Hawking-Gibbons temperature (de Sitter
temperature) given by the inverse of the Hubble scale, which is in turn related to the cos-
mological constant or vacuum energy of the metastring in that background. The vacuum
energy is, following the calculations from Section 2, bounded by the geometric mean of the
Hubble and the Planck scales.
Then one should examine a Gibbs-like ensemble of matrix-like variables at the de Sitter
temperature that describes the thermodynamics of de Sitter bound state. By construction
such an ensemble of matrix-like variables obeys the rules of the Voiculescu non-commutative
probability theory with infinite statistics. Then the desired relation between the variance of
energy and its average (and the square root number of degrees of freedom N) follows. Note,
that the modular Hamiltonian associated with the de Sitter horizon is the Hamiltonian of the
de Sitter bound state — so we also have the desired statement about the relation between
the modular Hamiltonian and its variance [82].
The spacetime “partons”, that is the elementary constituents, together with the back-
ground, from which the de Sitter bound state is made, should be associated with matrix
elements/entries, and the open strings between them. Those are the “spacetime atoms”
that obey infinite statistics. (As individual quanta, such as the partons of Matrix theory,
they are simply gravitons, with familiar spin-2 statistics. There is not contradiction here,
because, as reviewed in Section 2, EFT with an extra input from holography leads to the
same effective result for the vacuum energy as does a more fundamental description based
on modular regularization of phase space and its consequences, including infinite statistics.)
Thus non-perturbative metastring theory provides a concrete model of modular spacetime
and spacetime quanta obeying infinite statistics (consistent with the holographic bound),
and having the free energy of an effective two dimensional theory, that leads to the relation
between the fluctuation of the modular Hamiltonian and its expectation value, and further-
more, implies an effective collective field description of gravitons as spin-2 quanta of the
effective field theory of the gravitational field, as well as the unique temporal dependence of
the matrix model Green’s function [88] that could be experimentally tested.
31
4.4 Quantum Spacetime and Quantum Gravity Phenomenology
In this subsection we can pose the question of quantum gravity phenomenology associated
with the generic concept of modular spacetime and relate our discussion to [82] and the
subject of experimental probes of quantum spacetime based on gravitational interferometry.
We start by repeating the argument we presented for the vacuum energy [8, 29], as
reviewed in Section 2, which was based on the modular regularization of phase space and
the holographic/Bekenstein bound, albeit in the case of the causal diamond associated with
a gravitational wave detector (this set-up is observer dependent, and that fits our notion
of relative (observer-dependent) locality). In particular, we want to use Bekenstein’s bound
that is associated with the causal diamond of a gravitational interferometer. Then we obtain
that the characteristic scale of the vacuum (quantum spacetime) fluctuations is given by the
see-saw (ie, geometric mean) formula
p
δ ∼ l lP , (4.26)
where l is a characteristic length associated with the interferometer, and δ is the characteristic
length associated with the vacuum energy fluctuations due to the modular structure of
covariant phase space as in our recent papers [8, 29]. As observed by Zurek, δ/l is of the
order of the LIGO-VIRGO sensitivity (∼ 10−20 ) for l of order of a kilometer. By our
general argument, this could be a probe of modular spacetime and its modular structure,
in analogy with the Brownian-motion-like probes of the atomic structure of matter [96–99].
This reasoning makes sense, given Einstein’s basic relation hx2 i ∼ t, between the average
of the distance squared and the elapsed time, derived from the properties of the Gaussian
distribution associated with the relevant diffusion equation of the Brownian movement. In
our case, δ 2 ∼ hx2 i, and also l ∼ t. The Gaussian distribution encountered in the classic
analysis of Brownian motion is an example of a classical Boltzmann distribution. In our case
we are dealing with quanta of spacetime that obey the quantum analog of the Boltzmann
distribution which leads to the same relation between the fluctuation of energy and its
average, as derived in the previous subsection.
Note that in the case of the cosmological constant (cc) problem, l was identified with
the cosmological horizon (CH) (lCH ∼ 1027 m) which gave us the observed characteristic
scale associated with the cosmological constant (δcc ∼ 10−4 m). In the case of the gravita-
tional wave interferometer, like LIGO, l ∼ 103 m and thus the characteristic vacuum energy
scale in that case is δ ∼ 10−16 m. Also, the number of phase space cells in 4-dimensional
2
spacetime was Ncc ∼ lCH /lP2 ∼ 10124 , whereas in the case of the gravitational interferome-
ter N ∼ l2 /lP2 ∼ 1076 , which implies that per direction of spacetime we have N 1/4 ∼ 1019
“spacetime atoms” (which, as a historical remark, is not that different from the value of the
Avogadro number measured by Perrin following Einstein’s, Smoluchowski’s and Langevin’s
theoretical work [96–99]). In analogy with the classic treatment of Brownian motion [96–99]
one should be able to measure this effective “spacetime Avogadro number” N 1/4 ∼ 1019
using gravitational interferometry.
32
We note that the whole discussion presented in our recent paper on the vacuum energy
problem and gravitational entropy that was specifically devoted to 4 = 3 + 1 spacetime
dimensions can be formally extended to any D+1 spacetime dimensions. We just summarize
the relevant formulae (by setting all numerical factors to be of order one, for simplicity) [9]
lD−1 1 1
lD+1 ΛD+1 ∼ N, N∼ , ΛD+1 ∼ , λcc ∼ ΛD+1lPD−1 ∼ , (4.27)
lPD−1 l2 lPD−1 l2
(The first formula follows from the volume of phase space, the second from the holographic
bound in D + 1 spacetime dimensions, the third expression is the consequence of the first two
and the fourth expression is the definition of the cosmological constant in D + 1 spacetime
dimensions.) Thus the geometric mean relation from 4 spacetime dimensions is generalized
to Λ ∼ 1/lΛ
(D−1)/(D+1)
lΛ ∼ l2/(D+1) lP (4.28)
Indeed, for D = 3 we recover the seesaw/geometric mean formula cited in the beginning of
this subsection.
This general formula is not true for the situation considered in [82]: The geometric mean
formula between the fluctuation of energy and the energy mean (as derived in the previous
section) is valid in any number of dimensions and it is a square root/geometric mean/seesaw-
√
like formula δ ∼ l lP that we derived using the general properties of infinite (or quantum
Boltzmann, or quantum distinguishable) statistics. However, in 4 spacetime dimensions the
vacuum energy formula and the general formula derived from infinite statistics regarding
the relation between the energy fluctuation and its mean, coincide. Thus, gravitational
interferometers could be used as probes of quantum distinguishable statistics associated
with 4-dimensional spacetime quanta16 .
Moreover, as reviewed by Gross and Gopakumar [88] (see also [100]), the analog of the
Gaussian distribution for infinite statistics is the Wigner semicircle law for the eigenvalues
mi of an infinite square matrix M, so that the density ρ(m) of eigenvalues scales as
1√
ρ(m) = 4 − m2 , (4.29)
2π
R 1 2
for the Gaussian unitary ensemble D[M]e− 2 Tr[M ] . This is in turn represented by the
following operator expression in non-commutative probability theory
M = a + a† , (4.30)
with a and a† satisfying the free (Cuntz) algebra aa† = 1. The above semicircle distribution
for the eigenvalues of the spacetime coordinate operators should be observed in the exper-
iments that probe infinite statistics of spacetime quanta. This will be reflected in the time
dependence of the relevant Green’s function as implied by infinite statistics in the simplest
case of the Wigner semicircle distribution, as discussed by Gross and Gopakumar [88].
16
Another important smoking gun for the modular spacetime is associated with triple and higher order
interference phenomena as discussed in the context of the “gravitization of quantum theory” [54, 55, 72].
33
5 Quantum Gravity, UV/IR Mixing and Particle Physics
One of the main claims of this review is that quantum gravity effects can be already de-
tected in the context of the measurements of the cosmological constant, and that, viewed
as gravitized quantum theory, quantum gravity can be probed in experiments that check for
higher order (non-Born) quantum interference, as well as the probes of statistics of space-
time quanta using gravitational interferometry. The main point here is that quantum gravity
phenomenology occurs at scales far away from the Planck scale, and also in statistical, macro-
scopic settings. This is very much akin to the inherently quantum nature of macroscopic
phenomena of black body radiation, or the hardness of matter, or other various macroscopic
effects such as magnetism, conductivity, etc.
Most of these macroscopic quantum effects are based on quantum (Bose-Einstein or
Fermi-Dirac) statistics. As we argued in the previous section, in principle, spacetime quanta
have infinite (quantum distinguishable) statistics and they can be probed via the gravita-
tional analog of the Brownian motion experiment, but that probe is indirect. It is amusing
to speculate that spacetime quanta might be confined in analogy with quarks, and this ap-
pears quite natural from the point of view of the matrix model realization. One reason for
confinement of the spacetime quanta into what we identify as classical spacetime might be
via the crucial appearance of higher order quantum correlations (responsible for higher order
quantum interference) between these quanta. In analogy with other macroscopic quantum
phenomena one could think about all sorts of properties of spacetime and their origins in
quantum gravity: the origin of time17 from the rank and strength of the higher order correla-
tions, pointing to the purely quantum origin of time; similarly, the quantum origin of space
from the size of the higher order correlations; the origin of time arrow (and the impossibility
of closed timelike curves) from the second law of thermodynamics that involves higher order
correlations; the origin of causality (and thus, horizons) from the relation between the rank
and the size of higher order correlations; the origin of inertial frames from the “local” quan-
tum frames associated with “local” (in the space of states) quantum basis, and of inertia
as the macroscopic leftover of the gravitization of quantum theory and the relation between
short and long distances (or low rank and higher rank quantum correlations — a sort of
quantum Mach’s principle); the equivalence principle as the left over of the QG=GQ dictio-
nary (which could be understood as the quantum equivalence principle), etc. Of course, all
these possibilities are quite speculative at the moment.
In this section we claim that the already observed spectrum of masses of fundamental
particles is indicative of quantum gravity phenomenology at much lower scales than the
traditional Planck scale, MP ∼ 1019 GeV. Following the presentation in the recent paper [10],
we show that the above astro-particle conceptual arguments and ensuing calculations turn
out to have rather concrete consequences regarding the — well measured — (Standard Model)
17
In string theory the Liouville field, coming from the conformal anomaly (a quantum effect), and repre-
senting the conformal mode of the world-sheet 1 + 1 quantum gravity, appears as the natural clock, from the
rewriting of the critical string theory as a non-critical string in one extra, timelike, dimension [19].
34
particle physics. More precisely, we argue that the UV/IR mixing induces three separate
“intermediate” mass-scales (see also Figure 1): (1) the Higgs mass-scale, MH ∼ 102 GeV,
(2) the Bjorken-Zeldovich mass-scale, MBZ ∼ 7 MeV, (3) the mass-scale of the Standard
Model decoupling from dark matter, MSM ∼ 4.5×1014 GeV. In turn, these then induce the
detailed pattern of precisely three generations of Standard Model fermion masses and their
CKM/PMNS mixing. In particular, this enables us to write explicit predictions for the still
undetected neutrino masses.
Cosmological Scale: Reapplied within the metastring framework and its modular phase
space of the Heisenberg algebra (3.15) described in §§ 3.1 and 3.5, the above analysis of the
vacuum energy (including the Bekenstein bound) now produces a bound on the Higgs mass
and vev — since the latter specifies the vacuum.
In a 4+4-dimensional modular spacetime (X µ , X̃ ν ) with NΛ fluxes (see §§ 3.4–3.5) and
respective length scales lΛ and ˜l,
(lΛ ˜l )4 = NΛ (ℓ2s )4 (5.1)
is analogous to the relation (2.22), for the first of the Heisenberg algebras in (3.15). Together
with the relation NΛ = lΛ2 /lP2 , stemming from the holographic bound [16, 17] for the effective
spacetime associated with the vacuum energy, this produces
2 1/4 1/2
˜ 2 lΛ 2 lΛ
lΛ l = ℓ s 2 = ℓs . (5.2)
lP lP
The string and the Planck lengths, ℓs , lP , are related via the string coupling gs
g s ℓ s = lP , i.e. Ms = gs MP , (5.3)
with Ms and MP the corresponding mass scales, which sets the dual spacetime scale:
2 1/2
˜l = √ ℓs = lP lP . (5.4)
lΛ l P gs2 lΛ
18
Whereas the effective value of ℓs may well differ in 10 or any other spacetime dimension [101, 102], here
we focus on its 4-dimensional value.
35
Writing ˜l = η lP produces (see Table 1):
r r
1 l P 1 MΛ def
˜l
gs2 = = =⇒ gs < 1 when η = & 10−15.5 . (5.5)
η lΛ η MP lP
That is, gs < 1 and the one-loop computation of the (metastring) partition function is a
good approximation as long as ˜l . 10−21 m, and certainly so if the “dual (part of) phase
space” (defined by the dual spacetime and dual momenta) is of Planck size — which we
assume hereafter, for simplicity.
Higgs Mass: A formula for the Higgs mass was recently obtained (in bosonic string theory)
by Abel and Dienes [21],
2 MΛ4 gs2 Ms2
MH = ξ 2 − hX i , (5.6)
MP 8π 2
by relying primarily on modular invariance, which is however a universal stringy feature.
The ξ-term provides the modular completion to the second term, with a suitably normalized
insertion in the second moment of the partition function, X , for which hX i = −| hX i |. This
result in fact also follows from the foregoing analysis in the metastring framework: Indeed,
using (5.3) and (5.5) recasts (5.6) as
M 2 2 | hX i | p 2
Λ
MH2 = ξ + 2
MΛ MP , (5.7)
MP 8π
which is a simple numerical combination19 of the familiar “seesaw” (MΛ2 /MP ) and geometric
√
mean ( MΛ MP ) terms, both reflecting the seesaw relation of two scales, MΛ and MP .
The second term dominating the first one (∼ 10−34 eV), we have
r r r
| hX i | | hX i | p | hX i |
MH ∼ gs Ms 2
= gs2 MP 2
∼ MΛ MP , (5.8)
8π 8π 8π 2
recovering the string-theoretic seesaw formula for the Higgs mass [9], with the very realistic
numerical value as shown in Table 1. We emphasize that (5.8) is to be understood as a
bound on the Higgs mass, as is (2.29) for the cosmological constant.
Summary: Akin to our cosmological constant arguments and result (2.29), a stringy see-
saw formula (5.8) also follows for the Higgs mass: (1) within the metastring formulation of
string theory and its modular spacetime, by (2) combining the [x, x̃] 6= 0 non-commutativity
and holography in x-space, and by (3) assuming that x̃ is of the Planck length size. As with
the vacuum energy, (5.8) is also a bound provided by the size of the phase space and the
Bekenstein bound in which the effective length scale is associated with vacuum energy lΛ . In
this calculation, the two Heisenberg algebras in the metastring formulation ([x, p] and [x, x̃])
are mutually consistent.
19
| hX i | ∼ 10−1 is consistent with Abel & Dienes’ results [21].
36
Having applied this logic to the formula for the Higgs mass à la Abel & Dienes [21]
(derived in canonical bosonic string theory with manifest stringy modular invariance, but
also compatible with its metastring formulation), we have not only arrived at a stringy
bound for the Higgs mass, but also at a completely stringy view of the hierarchy problem.
So, both the Higgs mass and the hierarchy problem are direct consequences of the new view of
quantum gravity as gravitized quantum theory. The hierarchy problem is directly tied to the
vacuum energy problem, whereby the resolution of both lies in the fundamental (modular)
phase-space approach combined with a Bekenstein bound on the number of relevant degrees
of freedom. This new and unified understanding of these two central hierarchy problems
naturally points to metastring theory, and (as we outline in the next section) it can also
address the problem of fermion masses.
Let us however conclude this section by addressing the naturalness of the above values
for N, relevant for both hierarchy problems: the cosmological constant and the Higgs mass.
Both in statistical physics and in QFT, it is well known how to sum over contributions
of closed diagrams: Simple combinatorics ensures that this is an exponent of the partition
function associated with a closed loop (handle, for strings). As pointed out in Section 2.1,
the QFT vacuum partition function is Zvac = exp(ZS 1 ), with S 1 the circle of a vacuum
loop traced by a particle; in string theory, one just replaces S 1 → T 2 [8]. For the case
of dynamical Born geometry [53] (a generic feature of quantum gravity in the metastring
iS
formulation), the usual path integral measure eiS should be effectively replaced by ee after
summing over handles of a dynamical quantum geometry20 , where in the approximation of a
dilute gas of handles we have taken that the effective partition function is just the canonical
one. Summing over handles in the foamy quantum space [72], from the point of view of
the canonical complex geometry of quantum theory, thereby yields an effective action which
is essentially eiS . In the Euclidean formulation, this implies that the effective action at
some scale sensitive to gravity can be exponentially removed from the natural scale of Planck
gravity, indicating that the Higgs scale may well be where effects of quantum gravity could
be seen.21 Essentially, we claim the naturalness of the hierarchy of scales between the Higgs
and the Planck scale ultimately to be a quantum gravity effect, associated with “gravitizing
the quantum” [53–55]. Thus, the effective value of N (per spacetime direction) that features
in both hierarchy problems, the cosmological constant problem and the problem of the Higgs
mass, is indeed naturally expected to be of the order of the familiar Avogadro number, and
it is only genuine in the context of quantum gravity (or gravitized quantum theory) and
quantized spacetime.
20
It is tempting to associate this double exponential with another severe fine tuning problem in cosmology
— that of the initial state [9].
21
Indeed, see the large class of widely usable toy models [101–104].
37
5.2 Quantum Gravity and Fermion Masses: General Comments
The foregoing discussion shows that the same reasoning and computation successfully pro-
duces seesaw-like formulae, (2.29) for the cosmological constant, and (5.8) for the Higgs mass.
The weak string-coupling estimate (5.5) is consistent with the lowest-order perturbative com-
putations and analysis of the string partition function, and supports such expectations also
for the (meta)particle limit. Indeed, for both particles and strings, the cosmological constant
was thus shown to be bounded by the phase space volume, its modular regularization and
Bekenstein bound. In turn, Abel and Dienes’ stringy Higgs mass formula (5.7) [21] produces
the above, evidently analogous result — the logic of which is not attainable in EFT.
Both of these results follow from the new, “QG=GQ” view of quantum gravity as
gravitized quantum theory. The analogous solutions to these hallmark problems only have
differing contextual choices of the UV and IR scales: For both, MP is the natural UV scale.
On the other end, the Hubble/cosmological horizon provides a natural IR scale (MCH ) for
the cosmological constant scale, MΛ (2.29), which in turns is the IR scale for he Higgs mass,
MH in (5.8). The result of this reasoning is bound to agree with Abel and Dienes’s explicit
stringy result (5.7) since the vev of the Higgs field specifies both the Standard Model vacuum
as well as the Higgs mass.
In fact, since the same Higgs vev also provides masses to all charged Standard Model
fermions, the above reasoning should extend also to those. Whereas MH is also relevant for
neutrinos, the above reasoning will extend to them differently. However, before delving into
a derivation of the cascading seesaw formulae in Table 1, several frame-setting comments
and observations are in order.
Criticality: whereby the top-quark mass may be related to the Higgs mass as proposed
by Froggatt and Nielsen [105], is our first motivation. The running top and the Higgs mass
are related through the running top Yukawa coupling and the Higgs self coupling evaluated
at the running scale µ that is given by the top and Higgs masses. Given the top pole mass
mt , that is the physical mass of the top, and given the top running mass Mt (µ) [105]
mt 4 αs (mt ) αs (mt ) 2
=1+ + 10.95 , (5.9)
Mt (mt ) 3 π π
where αs (mt ) is the running strong coupling constant evaluated at the top mass. The Higgs
mass is already determined in the previous section, and we only have to set the running
scale µ at the first and the second minimum of the Higgs potential, which by criticality, are
given by the electroweak scale and the Planck scale, respectively [105]. The mass of the
top follows then from the mass of the Higgs. (One could repeat the same analysis for the
bottom quark and the tau lepton.) This, in turn, implies, via (5.7), that the mass of the
top also could be related to the cosmological constant — because the Higgs mass is. Again
by dimensional analysis, as in (5.7)–(5.8), the analogous fermionic formulae are expected to
be of the form mψ ∼ gs Ms , up to the multiplicative coefficients implied by stringy modular
38
invariance. This suggests a seesaw formula akin to the one for the Higgs mass (5.8), however
with appropriate UV- and IR-scales.22 The claim here is that such seesaw formulae relate
seemingly independent fermionic masses (in different generations) in the Standard Model.
In essence, this reasoning provides for the origin of different generations, starting from the
heaviest fermions, and predicts that there can exist no heavier generations of Standard Model
fermions.
To assess appropriate UV and IR scales for such fermionic seesaw formulae, recall that
the charged fermion masses (mt , mb and mτ ) are related via the RG equations for the heaviest
fermions in explicitly computable stringy models [106], and thus are natural candidates for
the UV scales. As to an appropriate IR scale, we present below an entropy argument that
leads to a scale attributed to QCD, but an order of magnitude smaller than the standard
ΛQCD ; we call this the Bjorken-Zeldovich scale, MBZ ≃ 7 MeV. This new low energy scale
(that is very close to the scale of Big Bang Nucleosynthesis) can be also viewed as another
fundamental infrared quantum gravity scale. We then find (as observed by Bjorken in a
completely different context [22]) that this MBZ can, with the masses of the heaviest charged
fermions as the UV scale, parametrize the masses of all remaining charged fermions.
This observation can be properly justified only by a computation of the bound of the
partition function of the Standard Model in the modular polarization, which by the already
explicit computation of the cosmological constant is given by the volume of phase space.
Relating then the number of phase space cells, in modular regularization, to the Bekenstein-
like bound with the UV scale given by the masses of the heaviest quarks and the heaviest
lepton then reproduces Bjorken’s expressions [22, 23].
Seesaw Structure: The foregoing discussion, including√the stringy result (5.7), involves
two types of formulae: The geometric mean, m < (m′ ∼ mM ) < M, is here implied by
the non-commutative, symplectic structure of Born geometry, ωab in (3.12). The standard
“seesaw,” (m′′ ∼ m2 /M) < m < M, is familiar from neutrino physics and is here of the
T-duality type, implied by the bi-orthogonal structure of Born geometry, ηab in (3.12). The
presence of the double metric, Hab in (3.12), is what allows the doubling of the heaviest
mass in the first place. This provides for three distinct masses and is, essentially, our key
observation here.
This dovetails with the fact that there are three generations, and meshes nicely with
the present experimental constraints on other generations of quarks and leptons. In what
follows, the bounds on the charged fermion masses take the form of these seesaw relations
(as used for the cosmological constant and also for the Higgs mass): with MU V identified
with the heaviest mass, the lighter copies are MIR -multiples of numerical factors that are
22
This indeed follows Weinberg’s general idea, “in some leading approximation the only quarks and leptons
with nonzero mass are those of the third generation, the tau, top, and bottom, with the other lepton and
quark masses arising from some sort of radiative correction” [6] — except, the lower fermion masses are here
generated by variants of the T-duality seesaw mechanism from a stringy non-perturbative effect; see § 3.5.
39
solely the square-root of ratios of the UV and IR scales, or the other way around:
r
MU V p
MIR = MIR MU V , for the middle, and (5.10a)
MIR
r s
MIR M2
IR
MIR = MIR , for the lightest. (5.10b)
MU V MU V
With the UV and IR scales as reasoned above, one expects the numerical factors in (5.10)
to be square-roots
p p Analogously, the dominant X -term in (5.7) gives MH ∼
of their ratios.
2
gs MP = MΛ /MP MP = MΛ MP /MΛ . Higher powers of these square-root factors then
correspond to higher powers of gs2, and are expected as (string-perturbative) corrections
to (5.7).23 By the same token, higher powers of the square-root factors in (5.10) are expected
as corrections of these formulae. For example, the standard seesaw-formula from the original,
neutrino physics,
MIR2 M r M 2
IR IR
= MIR = MIR , (5.11)
MU V MU V MU V
features the square of the numerical factor in (5.10b), and is expected to correspond to an
additive correction to (5.10b).
Also, let’s assume that a fermionic version of the stringy result (5.6) can be derived,
with a corresponding insertion vev, hXψ i, proportional to the gauge charges of the fermion
ψ as indeed is the case for the Higgs field [21]. Then: (1) for charged leptons, Xψ 6= 0, the
second term in a (5.6)-like formula again dominates, and formula (5.10a) follows. (2) For
chargeless neutrinos, Xψ = 0, only the first term in a (5.6)-like formula remains, and (5.11)
is the only contribution.
The remaining (T-duality type) seesaw formula (5.10b) stems from the central property
of the zero modes of the metastring captured by the action of the metaparticle (3.14), and
especially the constraint p·p̃ = µ. This is precisely the second, “seesaw-light” type relation,
2
where we identify µ = MBZ and the size of the dual momentum space with the relevant
charged fermion mass. Unlike the first seesaw formula (5.10a), which essentially follows from
the phase-space-like structure and so is associated with the symplectic form, this second
seesaw formula (5.10b) is induced by the bi-orthogonal structure of Born geometry.
Ideally, one would need a precise fermionic analogue of the Abel-Dienes formula for the
Higgs mass in string theory [21]. In the absence of such explicit formulae, we identify key
seesaw features that connect our approach to Bjorken’s observations [22, 23], which we then
also extend to the CKM matrix (like Bjorken), but also to neutrinos and the PMNS matrix
(in ways different from Bjorken). We find it intriguing that the same logic used for the
computation of the cosmological constant extends, first to the Higgs mass, and then also to
the masses of all quarks and leptons. In fact, this strongly suggests and a precise fermionic
analogue of the Abel-Dienes formula [21] must exist.
23
Also, the evident gs → gs−1 map between (5.10a) and (5.10b) would seem to indicate that S-duality must
be involved in an underlying stringy derivation of such formulae.
40
While these seesaw features do appear to be cohesive and coherent, a firm proof would
require the formulation of an explicit treatment of the Standard Model (SM) as a modu-
lar QFT : Every SM field φ is defined over both spacetime and the dual (momentum-like)
spacetime, φ(x, x̃), with an intrinsic non-commutativity [42, 44], [x, x̃] = iℓ2nc , where ℓnc is in
principle contextual, and not necessarily the string length or the Planck length.24 By con-
struction, such a formulation would have a natural solution of the vacuum energy problem,
and then, we conjecture, would also lead to the formulae for the fermionic masses presented
below. Such a modular SM would thereby imply relations between masses of different fermion
generations that are invisible to the standard QFT form of the SM. Such a modular QFT
form of the SM can be also embedded in the metastring, which suggest a completely new
(and complementary) view on the origin of the SM in string theory, as compared to the tra-
ditional one based on Calabi-Yau compactifications in the point-field limit QFT [19]. This
should indicate that there are missing concepts (modular spacetime, modular polarization,
Born geometry, modular fields, metaparticles and metastrings) in the usual approach, and
that the introduction of these missing concepts to the canonical approach would yield the
results discussed in this paper.
Unlike the very concrete foregoing statements about the vacuum energy problem and
the problem of the Higgs mass, our present discussion of fermion masses is just a working
conjecture at the moment. We now turn to the implementation of this general set-up by
following our recent presentation [10, 11].
The Bjorken-Zeldovich
p scale: The relation (2.25) specified the “spacetime Avogadro
1/4
number,” N ∼ l/lP ∼ 1031 . However, in 3-dimensional space and as expected from
extensive non-gravitational entropy, N defines a length-scale, lBZ (see below for naming):
def def (2.29)
l3 /lBZ
3
∼ N ∼ l2 /lP2 , ⇒ 3
lBZ ∼ l3 /N ∼ l lP2 ∼ lcc
2
lP , i.e., 3
MBZ ∼ MΛ2 MP . (5.12)
More precisely, the lBZ and MBZ scales have been deduced from: (1) our N (2.25), (2) the
Bekenstein bound for gravitational degrees of freedom, (3) the fact that matter and spacetime
degrees of freedom are “two sides of the same coin” in (meta)string theory, and (4) the
extensive nature of entropy for the matter degrees of freedom.
The numerical estimate in Table 1, MBZ ≃ 7.2 MeV, turns out to be exactly the value
used by Bjorken [22]25 to parametrize the observed masses of Standard Model quarks and lep-
tons, which is why we call MBZ the “Bjorken-Zeldovich scale.” The continued relations (5.12)
24
Both of these scales, ℓs and lP , turned up naturally in the discussion in Section 3.5, but note that the
effective, physically relevant 4-dimensional Planck scale may be removed, even exponentially much, from the
underlying fundamental scale, e.g., in the large class of models discussed in [101–104].
25
Bjorken seems to have been inspired by the work of the Oxford group [107], and discussed MBZ in a
radically different context of the MacDowell–Mansouri approach to gravity, and in particular, the Friedmann-
Robertson-Walker cosmology in that formulation [22].
41
express it in terms of the cosmological horizon and the Planck scale, and provides a deriva-
tion that is, as best as we know, completely new. These three key scales, the cosmological
constant mass scale, MΛ , the Higgs mass scale, MH , as well as the Bjorken-Zeldovich scale,
MBZ , are thereby all ultimately determined in terms of the Hubble (MCH ) and Planck mass
(MP ) scales; see Table 1 and Figure 1.
There is an another, exceedingly curious relation, and independently corroborating the
preceding derivation: The seesaw expression with the proton mass mp ,
mp 2 (938.27 MeV)2
= ≃ 7.0288 MeV (5.13)
MH 125.25 GeV
almost exactly reproduces the above-derived mBZ (5.12)! Furthermore, the structure of the
seesaw formula is itself aligned with our observations throughout this report: The well un-
derstood electroweak phase transition (i.e., the Higgs mass, derived in § 5.1, above) sets the
UV scale (in the denominator) of (5.13). In turn, the numerator specifies an appropriate IR
mass-scale, well known to be associated with QCD: it is the mass of the lightest and only sta-
ble QCD bound state, the proton.26 This then identifies the mass-scale (5.13), equal to MBZ ,
as characteristic for charged SM-fermions: The QCD-characteristic (5.13) is evidently rele-
vant for quarks, and then also to electrically charged leptons, via EM-radiative corrections.
Analogous radiative corrections to neutrino masses are below MH suppressed by weak gauge
boson masses, and will have to be determined by different mass-scales; see (5.22)–(5.24), be-
low. Let us also add that the oft-cited and ∼ 20 times larger ΛQCD ∼ 150 MeV is defined by
the Landau pole, i.e., the momentum transfer at which the (perturbatively computed ) strong
interaction coupling diverges; ΛQCD is thus a characteristic of the perturbative description
of QCD, rather than of QCD itself.
Quarks: Following the above reasoning, we start with the masses of the heaviest charged
SM fermions, mt , mb , mτ , as essentially being determined by the electroweak phrase transi-
tion, i.e., the Higgs mass. To this end, we write (see Table 1)
def def def
mt = Yt MH , mb = Yb MH , and mτ = Yτ MH , (5.14)
7 Yt Yt
were Yt ≈ /5 , Yb ≈ /42 , and Yτ ≈ /100 , (5.15)
are the concrete numerical values of these Yukawa couplings; see below for a justification
of these estimates. (For a concrete computation of these masses in a string theory model,
see [106].) These masses serve as analogs of the UV scale in our Higgs mass formula (5.8),
whereas the Bjorken-Zeldovich scale, MBZ ≈ 7 MeV, acts as the natural IR scale. Note
that the top mass mt is essentially tied to the Higgs scale,27 which in turn is given by the
26
The commonly cited Landau pole, ΛQCD ∼ 100 MeV, is the momentum exchange magnitude where
the perturbatively computed coupling parameter diverges, and so seems to be a hallmark of a perturbative
description rather than an intrinsic characteristic of QCD itself. Also, likeliest candidates for pure-glue
bound states have masses above 2 GeV [108].
27
To this end, we cite the well-known argument based on criticality of the Standard Model that relates the
masses of the top quark and the Higgs boson [105] (see also [109, 110], for landscape-motivated discussions).
42
(geometric mean) seesaw formula of the vacuum energy scale and the Planck scale. Thereby,
the top quark mass is ultimately also given in terms of the Hubble and Planck mass scales.
Analogously to (5.8) for the Higgs mass, the (geometric mean) seesaw relation then produces
the charm mass in terms of MBZ and mt (cf. the observed value in parentheses [24]):
p r
mt
mc ∼ MBZ mt = MBZ ∼ 1.10 (1.27) GeV. (5.16)
MBZ
Next, using the bottom-quark mass scale28 (instead of mt ) and the same Bjorken-Zeldovich
scale as the characteristic vacuum energy scale of matter, the same seesaw relation yields
the mass of the strange quark
p r
mb
ms ∼ MBZ mb = MBZ ∼ 171 (93.4) MeV. (5.17)
MBZ
Bjorken estimates the up- and down-quark masses essentially at the Bjorken-Zeldovich scale:
mu ∼ MBZ and md ∼ MBZ , but models the actual relation md > mu with ad hoc factors [22].
Independently, the masses of the lightest quarks may be deduced from chiral perturbation
theory as mu ∼ 2 MeV, md ∼ 5 MeV. However, apart from non-commutativity that led
to (5.16) and (5.17), our seesaw structure reasoning above involves also the inherent metas-
tring/metaparticle T-duality, which induces the familiar “seesaw-light” relation. This then
leads to the following estimates (actual values in parentheses [24])
r
2 MBZ
mu ∼ MBZ /mc ∼ MBZ ∼ 10−2 MBZ ∼ 10−1 (2.16) MeV. (5.18)
mt
This estimate turns out too small (by a factor of about 50), but is (importantly!) smaller
than the down quark mass estimate (also too small by a factor of about 16),
r
2 MBZ
md ∼ MBZ /ms ∼ MBZ ∼ 10−1 MBZ ∼ 1 (4.67) MeV. (5.19)
mb
The above reasoning thus automatically reproduces the 1st generation “mass inversion”:
(5.16) > (5.17) but (5.18) < (5.19), which is necessary for the proton to be stable while the
neutron decays. Thus, given the heaviest, top and the bottom quark masses, the two distinct
seesaw type formulae (non-commutativity and T-duality) produce quite realistic estimates
for the masses of the middle and the lightest quark generations.
Charged leptons: Turning to the charged leptons, the evident analogue of the top-quark
is the tau-lepton. From a naive stability analysis of the tau analogue of the hydrogen atom,
the mass of the tau is expected to be of the order of the mass of the nucleus, i.e. a GeV. This
is supported since the masses of the top, bottom quark and the tau lepton are all related by
28
For example, explicit calculation in the stringy calculation [106] ties, via RG equations, the mass of the
top quark to the mass of the bottom quark and the tau lepton, and so are all ultimately determined by the
Hubble and Planck mass scales.
43
the RG equations, as in the calculation of [106]. With the tau mass as given (again, from the
calculation of [106], and ultimately related to the Hubble and Planck mass scales, much as
the top and bottom quark masses are), the (geometric mean) seesaw estimate of the muon
mass is (actual value in parentheses [24]):
p r
mτ
mµ ∼ MBZ mτ = MBZ ∼ 112 (106) MeV. (5.20)
MBZ
Just as with quarks, the second (T-duality kind) seesaw relation then yields the electron
mass, given the calculated muon mass
2
r
MBZ MBZ
me ∼ ∼ MBZ ∼ 464 (511) keV. (5.21)
mµ mτ
This proposal thus reproduces 3 generations of charged Standard Model fermions and
their masses, by the framework of the dual space, the modular spacetime Born geometry, and
ultimately the metastring, i.e., by the intrinsic non-commutativity and covariant T-duality
of the metastring. The masses of the two lighter generations are induced from the masses
of the heaviest quarks and leptons, and are fixed by non-commutativity and T-duality, in
analogy with the reasoning that gives the Higgs mass and the cosmological constant. All of
these formulae are seesaw-like and contextual bounds. All of them ultimately reduce to the
IR size of the universe and the UV Planck length.
where the SM scale MSM is given by a “would-be unification scale” of the SM couplings (as
indicated by RG equations), ∼ 1015−16 GeV, and MH is the Higgs scale of around 1 TeV.
This Standard Model scale can be also related to the Hubble and the Planck scales MSM =
1/14 13/14
MCH MP ∼ 4.5×1014 GeV, as indicated in Table 1. This MSM also appears in Vafa’s
analysis [112], postulated as the scale at which decaying dark matter “gives a small kick to
its decay products.” The fully T-dual, chirally doubled description of stringy spacetime [5],
dark and visible matter mix via a Berry phase-like correlation term [103]. The mass-scale of
this correlation term defines the decoupling energy, below which the dark and visible matter
interact as required, only gravitationally. MSM ≃ 4.5×1014 GeV is thus naturally identifiable
with this dark-matter/visible-matter decoupling scale.
The mass of the heaviest neutrino (5.22) would then also be ultimately given in terms
of the Hubble and the Planck mass scales. The middle (“muon”) neutrino mass is then
given by a (geometric mean) seesaw formula, involving a low vacuum energy scale. Unlike
all quarks and charged leptons, neutrinos do not get their masses from the Higgs mechanism,
44
so the vacuum scale cannot be MBZ (characteristic for quarks and charged fermions) and
so must be the only other vacuum scale: the cosmological vacuum scale associated with the
cosmological constant (2.29):
p r
m3
m2 ∼ MΛ m3 = MΛ ∼ (10−2 − 10−2.5) eV. (5.23)
MΛ
By comparison, a similar mass value has been argued [113] to be natural by examining a
dimension 6 analogue of Weinberg’s operator, where a neutrino could acquires its mass from
a fermionic condensate controlled by the Bjorken-Zeldovich scale, with the electroweak cutoff
3
scale: m2 ∼ MBZ /m2H ∼ MΛ ; see (5.8) and (5.12).
Finally, the lightest (“electron”) neutrino mass is then estimated by the (T-duality)
seesaw formula r
2 MΛ
m1 ∼ MΛ /m2 ∼ MΛ ∼ 10−4 eV. (5.24)
m3
According to the Particle Data Group [24], the sum of neutrino masses (coming from cos-
mology) is bounded by 10−1 eV, which is satisfied by the above normal hierarchy of neu-
trino masses. Incidentally, the above reasoning predicts the normal hierarchy, i.e., that
(m3 /m2 ) < (m2 /m1 ). Also, these values satisfy the constraint on the square of the differ-
ences of masses, (10−2 –10−5 ) eV2 , coming from neutrino oscillation experiments. It would
exciting to learn if these values for the neutrino masses are experimentally falsifiable.
All these estimates for quark lepton and Higgs masses and for the cosmological constant
mass scale are upper bounds; this bound for mu and md essentially being given by MBZ . We
thus expect an attractor mechanism (as in [114]) that would “glue” all these values to their
upper bounds. This would be consistent with the existence of a moduli-free self-dual fixed
point in metastring theory [5] that could explain the apparent criticality of the Standard
Model parameters [105]. Finally, all these bounds on the fermion masses, much as the bounds
on the cosmological constant and the Higgs mass, are determined in terms of the Hubble
and the Planck mass scales.
The PMNS Matrix: Neutrino mixing is parametrized much as the CKM matrix, (5.27)–
(5.29), but using MΛ instead of MBZ as motivated in the above discussion of neutrino masses,
as well as by replacing mb → m3 , ms → m2 and md → m1 , as should be evident. We also
take into account that m3 is known up to a factor of 1/10 in the above formula for the
heaviest neutrino mass (the observed data [24, 25] is included in parentheses):
r r
MΛ MΛ MΛ
|Uµ3 | ∼ √ ∼ ∼ 0.50, (0.63) (5.30)
m3 m1 m3 m1
(5.22)
(essentially, (MΛ /m3 ∼ MSM /MP )1/4 ) as well as
r r
MΛ MΛ MΛ
|Uτ 1 | ∼ √ ∼ ∼ 0.13, (0.26) (5.31)
m3 m2 m3 m2
47
dynamical. This approach is based on the new calculation of the cosmological constant, Λcc
based on the phase space reasoning and the holographic bound. Gravitization of quantum
theory can be experimentally probed by searching for higher order quantum interference
phenomena in the presence of gravity, as well as by probing the statistics of spacetime
quanta via gravitational interferometry. By extending the computation of the cosmological
constant, we have also evaluated the masses of the Higgs boson, as well as quarks and leptons
and their mixing matrices.
Perhaps the most dramatic prediction of dynamical Born geometry implies “gravitiza-
tion of quantum theory” [53–55],and the presence of intrinsic and irreducible triple (and
higher order) interference [74] in the presence of gravity [72]. This would be a new quantum
probe of quantum spacetime and a new avenue in quantum gravity phenomenology [117,118].
√
Similarly, the crucial seesaw formula, δ ∼ lIR lU V , (with a characteristic IR length-scale
lIR and the characteristic UV length-scale lU V ) found in the context of the computation of
the vacuum energy, appears in other related contexts, such as the gravitational wave inter-
ferometry probes of quantum gravity; see, for example, [119]. In that context, our vacuum
energy calculation can be performed on the level of the causal diamond of the interferometer
(lIR being given by the length of the interferometer and lU V by the Planck length), leading
to the same seesaw formula, except interpreted as an empirical probe of modular spacetime.
As we emphasized in Section 4, the existing LIGO-VIRGO sensitivity is enough to test this
seesaw relation for interferometers with l of the order of a kilometer.
The geometric mean, seesaw formula for the associated length scale of the cosmologi-
cal constant, lcc (2.29), exhibits UV/IR mixing, and that makes lcc radiatively stable and
natural. The logic of this resolution of the cosmological constant problem, with input from
the Abel-Dienes stringy calculation, extends naturally to the Higgs mass (5.8). The same
idea applies to the masses and mixing of quarks and leptons. We have emphasized the
quantum gravity nature of the cosmological constant, the Higgs mass and the masses and
mixing angles of quarks and leptons. One important new ingredient in this reasoning is
quantum contextuality (instead of the standard anthropic reasoning) which stems from the
string/modular QFT vacuum being governed by Born geometry based on the modular phase
space view of quantum spacetime à la [8]. The crucial interplay of (1) phase space, (2) Born
geometry, (3) the Bekenstein bound, (4) mixing between ultraviolet (UV) and infrared (IR)
physics, and (5) modular invariance in string theory (in its intrinsically non-commutative,
metastring formulation) was emphasized throughout this review. All these are the essential
features of the new view of quantum gravity as gravitized quantum (or metaquantum) theory.
Throughout the review we have repeatedly stressed the purely stringy or quantum-
gravity-related effects which are fundamentally rooted in the properties of quantum space-
time. Such effects are not part of the usual EFT lore, largely because EFT is defined in
classical spacetime as a background. This might sound disturbing given the success of EFT.
Consequently, we have argued that EFT results, dressed up with holography, can be recov-
ered in a singular limit of our computation of the vacuum energy. Given the fact that the
usual compactification approach to string theory, and the associated string landscape and
48
swampland [9,120,121], are closely tied to EFT, we conjecture that the application of holog-
raphy in that context, and a seesaw relation between what are usually considered UV and IR
cut-offs in EFT [34], could lead to a top-down realization of our computations and results,
at a critical self-dual point (without moduli) which would hide the fundamental aspects of
our discussion: modular spacetime, Born geometry and the metastring formulation.
The four-dimensional nature of our discussion may in this approach be related to the fun-
damental properties of strings at high temperature in the early universe [51]. This could be
then generalized to the computations of the Higgs mass and the masses and mixing matrices
of quarks and leptons, as discussed in this paper, revealing, perhaps, an attractor mechanism
in string landscape and swampland. Similarly the observed cosmological background (de Sit-
ter spacetime) which is usually problematic in string theory [9] (and thus leads to various
conjectures in the landscape [120] and swampland approaches [121]) could be understood as
the natural cosmological bound state of the non-perturbative matrix model formulation of
metastring theory, and as such is a consequence of the view on quantum gravity as gravitized
quantum theory.
In conclusion, we list some further phenomenological implications of our work.
Our calculation of the cosmological constant introduces a new quantum number (2.21),
N, which may be probed in gravitational waves, via gravitational wave “echoes”: In par-
ticular, our result (2.25) relates the number of phase space boxes to the Bekenstein bound,
N ∼ l2 /lP2 . It can therefore be used for black holes, where l → lbh is the size of the black hole
horizon, where it is naturally related [122]. In this case, the relevant quantization number,
2
Nbh ∼ lbh /lP2 , for black holes is of the order of 1080 , and a possible observable feature of this
2
quantization, lbh ∼ NlP2 , might be via the “gravitational wave echoes” [123, 124] — in the
“quantum chaos” phase, given the enormous value of N. We also observe that in the context
of the non-perturbative formulation of metastring theory via a gravitized matrix quantum
theory discussed in section 3, black holes appear as natural astrophysical bound states of
the fundamental partonic quantum spacetime degrees of freedom. Various observational
astrophysical consequences of this picture are yet to be explored.
Furthermore, seesaw formulae for the SM fermion masses follow from the same reasoning
that lead to the cosmological constant (2.29) and the Higgs mass (5.8) seesaw formulae. In
that situation, a new Bjorken-Zeldovich scale can be deduced (by analogous reasoning) which
enters into Bjorken-like seesaw formulae for all masses of charged elementary fermions.30
This approach seems to proffer a new view on the observed three generations of quarks
and leptons as well as their respective mixing matrices. Here we point out an analogy with
critical phenomena and the mean field/Landau-Ginzburg (LG) approach which gives “square
root type” formulae, or the critical index of 1/2, without any anomalous dimensions, and
30
In a fermionic (5.7)-like formula, the Xψ -insertion term must be proportional to gauge charges, and so
is absent for neutrinos. In any explicit model-dependent calculation such as [106], the RG equations “tie”
the heaviest charged fermion masses to the electroweak scale, while for neutrinos the relevant RG equations
extend the UV scale to MSM ∼ 1015−16 GeV.
49
which are, in turn, introduced by a more precise renormalization group (RG) treatment
of the LG like description. In our case, the analogue of LG is the modular field theory
extension of the SM and gravity. Our formulae should therefore be understood in the “mean
field theory sense”. In the context of modular field theory (a consistent limit of metastring
theory) we also expect a double RG that is sensitive both to UV and IR scales [44]. Also, all
these formulae can be rewritten ultimately using only the Hubble (IR) scale MCH and the
Planck (UV) scale MP . These are the only two scales that appear in all expressions for the
cosmological constant, the Higgs mass and the masses and mixing of quarks and leptons.
We also point out that in the visible sector we ultimately have to work with modular
fields φ(x, x̃) [41]. This is not so in the Standard Model (SM) as it is understood at the
moment, but is implied by the modular polarization and our argument about the bounds of
fermion masses. Thus, the modular SM fields should know about the symplectic and also
the biorthogonal structures associated with x and x̃. (This suggests a kind of generalized
mirror symmetry in the visible sector.) This is what induces two distinct seesaw formulae
(one non-commutative/symplectic, and one T-dual/biorthogonal), naturally yielding three
generations in x-spacetime (a heavy fermion and its two seesaw copies). The invisible (dark)
sector is spanned by the dual fields φ̃(x, x̃), which may well be subject to a third quanti-
zation indeterminacy because of an induced non-commutativity between visible φ and dual
(invisible/dark) φ̃ fields (fuzzy dark matter). Thereby, while one may be able to deduce the
bounds on the parameters of the Standard Model (SM) in string theory/quantum gravity,
the ensuing indeterminacy in the parameters of the dual Standard Model (the dark sector)
should then be reciprocal to the relatively high precision (small indeterminacy) of the SM
parameters. (This would be in the spirit of the old “third quantization” proposal [125]).
We stress that metaparticles [42] (zero modes of the metastring) represent a generic
prediction of metastring theory and the dark matter sector can be seen as coming from a
dual Standard Model with a dynamics that is entangled/correlated with the visible Standard
Model [49]. The dark matter degrees of freedom are thus tied to the dual particles to the
visible SM particles [9]. Furthermore, this approach shows dark energy (modeled as the
cosmological constant) to be the curvature of the dual spacetime, and naturally small [9].
The natural relation between the dark matter and dark energy sectors in our formulation
(see also [48]), as well as the relation between the visible and dark sectors, offers, apart from
quantum contextuality, a new view on the coincidence problem in cosmology [120].
References
[1] D. Oriti, ed., Approaches to Quantum Gravity: Toward a New Understanding of Space,
Time and Matter. EBL-Schweitzer. Cambridge University Press, 2009.
[2] J. de Boer et al., “Frontiers of Quantum Gravity: shared challenges, converging directions,”
arXiv:2207.10618 [hep-th].
[3] Supernova Search Team Collaboration, A. G. Riess et al., “Observational evidence from
supernovae for an accelerating universe and a cosmological constant,”
Astron. J. 116 (1998) 1009–1038, arXiv:astro-ph/9805201.
[5] L. Freidel, R. G. Leigh, and D. Minic, “Metastring Theory and Modular Space-time,”
J. High Energ. Phys. 06 (2015) 006, arXiv:1502.08005 [hep-th].
[7] S. Weinberg, “The Cosmological Constant Problem,” Rev. Mod. Phys. 61 (1989) 1–23.
[8] L. Freidel, J. Kowalski-Glikman, R. G. Leigh, and D. Minic, “The Vacuum Energy Density
and Gravitational Entropy,” arXiv:2212.00901 [hep-th].
[9] P. Berglund, T. Hübsch, and D. Minic, “On de Sitter Spacetime and String Theory,”
arXiv:2212.06086 [hep-th].
[10] P. Berglund, T. Hübsch, and D. Minic, “String theory bounds on the cosmological constant,
the Higgs mass, and the quark and lepton masses,” Symmetry 15 no. 9, (8, 2023) 1660,
arXiv:2307.16712 [hep-th].
[11] D. Minic, “The vacuum energy problem in quantum gravity and the masses of elementary
particles,” Bulgarian J. Phys. 50 (May, 2023) 1–12, arXiv:2305.12593 [hep-th].
[13] L. Freidel, R. G. Leigh, and D. Minic, “Born Reciprocity in String Theory and the Nature
of Spacetime,” Phys. Lett. B730 (2014) 302–306, arXiv:1307.7080 [hep-th].
51
[14] P. Berglund, T. Hübsch, and D. Minic, “Mirror Symmetry, Born Geometry and String
Theory,” in Nankai Symposium on Mathematical Dialogues: In celebration of S.-S. Chern’s
110th anniversary, Y.-H. He, M.-L. Ge, C.-M. Bai, J. Bao, and E. Hirst, eds. Springer
Verlag, Singapore, Sept., 2022. arXiv:2111.14205 [hep-th].
[15] J. D. Bekenstein, “A Universal Upper Bound on the Entropy to Energy Ratio for Bounded
Systems,” Phys. Rev. D 23 (1981) 287.
[16] G. ’t Hooft, “Dimensional reduction in quantum gravity,” Conf. Proc. C 930308 (1993)
284–296, arXiv:gr-qc/9310026.
[17] L. Susskind, “The World as a hologram,” J. Math. Phys. 36 (1995) 6377–6396,
arXiv:hep-th/9409089.
[18] W. Fischler and L. Susskind, “Holography and cosmology,” arXiv:hep-th/9806039.
[19] J. Polchinski, String theory. Vol. 1: An introduction to the bosonic string. Cambridge
Monographs on Mathematical Physics. Cambridge University Press, Dec., 2007.
[20] J. Polchinski, String theory. Vol. 2: Superstring theory and beyond. Cambridge Monographs
on Mathematical Physics. Cambridge University Press, Dec., 2007.
[21] S. Abel and K. R. Dienes, “Calculating the Higgs mass in string theory,”
Phys. Rev. D 104 no. 12, (2021) 126032, arXiv:2106.04622 [hep-th].
[22] J. Bjorken, “Darkness: What comprises empty space?”
Annalen der Physik 525 no. 5, (2013) A67–A79.
[23] J. D. Bjorken, “Masses and mixings of quarks and leptons.” Slides, last checked: Nov., 2022.
https://nebula.wsimg.com/44afe7e009f4c9854e6dc1b8887bdbc8?AccessKeyId=D5AC63041E00FF1ED0E8
[24] R. L. Workman et al. (Particle Data Group), “Review of Particle Physics,”
Progress of Theoretical and Experimental Physics 2022 no. 8, (08, 2022) 083C01.
https://doi.org/10.1093/ptep/ptac097.
[25] I. Esteban, M. C. Gonzalez-Garcia, M. Maltoni, T. Schwetz, and A. Zhou, “The fate of
hints: updated global analysis of three-flavor neutrino oscillations,”
J. High Energy Phys. 2020 no. 9, (2020) 178.
[26] M. B. Green, J. H. Schwarz, and E. Witten, Superstring Theory, vol. 2: (Loop Amplitudes,
Anomalies and Phenomenology) of Cambridge Monographs on Mathematical Physics.
Cambridge University Press, Cambridge, 1987.
[27] J. F. Donoghue, “Cosmological constant and the use of cutoffs,”
Phys. Rev. D 104 no. 4, (2021) 045005, arXiv:2009.00728 [hep-th].
[28] J. Polchinski, “Evaluation of the One Loop String Path Integral,”
Commun. Math. Phys. 104 (1986) 37.
[29] L. Freidel, J. Kowalski-Glikman, R. G. Leigh, and D. Minic, “On the Inevitable Lightness of
Vacuum,” arXiv:2303.17495 [hep-th].
[30] L. Freidel, R. G. Leigh, and D. Minic, “Modular spacetime,”
Int. J. Mod. Phys. D24 no. 12, (2015) 1544028.
[31] L. Freidel, R. G. Leigh, and D. Minic, “Quantum Spaces are Modular,”
Phys. Rev. D94 no. 10, (2016) 104052, arXiv:1606.01829 [hep-th].
52
[32] S. Weinberg, “Problems in Gauge Field Theories,” in 17th International Conference on
High-Energy Physics, pp. III.59–65. 1974.
[34] A. G. Cohen, D. B. Kaplan, and A. E. Nelson, “Effective field theory, black holes, and the
cosmological constant,” Phys. Rev. Lett. 82 (1999) 4971–4974, arXiv:hep-th/9803132.
[35] S. Weinberg, The Quantum Theory of Fields, vol. 1: Foundations. Cambridge University
Press, 1995.
[36] A. Zee, Quantum Field Theory in a Nutshell. Princeton University Press, 2nd ed., 2010.
[38] D. Finkelstein, Quantum Relativity: A Synthesis of the Ideas of Einstein and Heisenberg.
Theoretical and Mathematical Physics. Springer Berlin Heidelberg, 2012.
[39] Y. Aharonov and D. Rohrlich, Quantum Paradoxes: Quantum Theory for the Perplexed.
Wiley-VCH, 2005.
[41] L. Freidel, R. G. Leigh, and D. Minic, “Noncommutativity of closed string zero modes,”
Phys. Rev. D96 no. 6, (2017) 066003, arXiv:1707.00312 [hep-th].
[44] L. Freidel, R. G. Leigh, and D. Minic, “Modular Spacetime and Metastring Theory,”
J. Phys. Conf. Ser. 804 no. 1, (2017) 012032.
[46] E. Barnes, J. J. Heremans, and D. Minic, “Non-Fermi Liquids, Strange Metals and
Quasi-metaparticles,” arXiv:2111.10479 [cond-mat.str-el].
[47] D. Edmonds, D. Farrah, D. Minic, Y. J. Ng, and T. Takeuchi, “Modified Dark Matter:
Relating Dark Energy, Dark Matter and Baryonic Matter,”
Int. J. Mod. Phys. D27 no. 02, (2017) 1830001, arXiv:1709.04388 [astro-ph.CO].
[48] D. Edmonds, J. Erlich, D. Minic, and T. Takeuchi, “Universal acceleration and fuzzy dark
matter,” arXiv:2405.08744 [astro-ph.CO].
[49] P. Berglund, T. Hübsch, and D. Minic, “String theory, the dark sector and the hierarchy
problem,” LHEP 2021 (2021) 186, arXiv:2010.15610 [hep-th].
http://journals.andromedapublisher.com/index.php/LHEP/article/view/186.
53
[51] R. H. Brandenberger and C. Vafa, “Superstrings in the Early Universe,”
Nucl. Phys. B316 (1989) 391–410.
[52] R. Penrose, “Palatial twistor theory and the twistor googly problem,”
Phil. Trans. Roy. Soc. Lond. A 373 (2015) 20140237.
[53] L. Freidel, R. G. Leigh, and D. Minic, “Quantum Gravity, Dynamical Phase Space and
String Theory,” Int. J. Mod. Phys. D23 no. 12, (2014) 1442006,
arXiv:1405.3949 [hep-th].
[56] A. Strominger, “Lectures on the Infrared Structure of Gravity and Gauge Theory,”
arXiv:1703.05448 [hep-th].
[57] L. Freidel, M. Geiller, and D. Pranzetti, “Edge modes of gravity. Part I. Corner potentials
and charges,” JHEP 11 (2020) 026, arXiv:2006.12527 [hep-th].
[58] L. Freidel, M. Geiller, and D. Pranzetti, “Edge modes of gravity. Part II. Corner metric and
Lorentz charges,” JHEP 11 (2020) 027, arXiv:2007.03563 [hep-th].
[60] J. J. Atick and E. Witten, “The Hagedorn Transition and the Number of Degrees of
Freedom of String Theory,” Nucl. Phys. B 310 (1988) 291–334.
[63] D. Minic and H. C. Tze, “Background independent quantum mechanics and gravity,”
Phys. Rev. D 68 (2003) 061501, arXiv:hep-th/0305193.
[65] D. Minic and H. C. Tze, “What is quantum theory of gravity?,” in 3rd International
Symposium on Quantum Theory and Symmetries, pp. 159–166. 1, 2004.
arXiv:hep-th/0401028.
54
[67] A. Ashtekar and T. A. Schilling, “Geometrical formulation of quantum mechanics,”
arXiv:gr-qc/9706069.
[70] Y. Aharonov and J. Anandan, “Phase change during a cyclic quantum evolution,”
Phys. Rev. Lett. 58 (Apr, 1987) 1593–1596.
[71] E. Witten, “Quantization of Chern-Simons Gauge Theory With Complex Gauge Group,”
Commun. Math. Phys. 137 (1991) 29–66.
[75] D. Minic, T. Takeuchi, and C. H. Tze, “Interference and Oscillation in Nambu Quantum
Mechanics,” arXiv:2012.06583 [hep-ph].
[77] H. Awata, M. Li, D. Minic, and T. Yoneya, “On the quantization of Nambu brackets,”
JHEP 02 (2001) 013, arXiv:hep-th/9906248.
[82] E. Verlinde and K. M. Zurek, “Spacetime Fluctuations in AdS/CFT,” JHEP 04 (2020) 209,
arXiv:1911.02018 [hep-th].
[83] K. M. Zurek, “On vacuum fluctuations in quantum gravity and interferometer arm
fluctuations,” Phys. Lett. B 826 (2022) 136910, arXiv:2012.05870 [hep-th].
[84] D. Minic, “Infinite statistics and black holes in matrix theory,” arXiv:hep-th/9712202.
55
[85] C. M. Ho, D. Minic, and Y. J. Ng, “Dark Matter, Infinite Statistics and Quantum Gravity,”
Phys. Rev. D85 (2012) 104033, arXiv:1201.2365 [hep-th].
[86] V. Jejjala, M. J. Kavic, D. Minic, and T. Takeuchi, “Dynamical dark energy and infinite
statistics,” Int. J. Mod. Phys. A 37 no. 17, (2022) 2242001, arXiv:2011.08852 [hep-th].
[87] D. Voiculescu, K. Dykema, and A. Nica, Free Random Variables, vol. 1 of CRM monograph
series. American Mathematical Society, 1992.
[89] A. Strominger, “Black hole statistics,” Phys. Rev. Lett. 71 (1993) 3397–3400,
arXiv:hep-th/9307059.
[90] O. W. Greenberg, “Example of infinite statistics,” Phys. Rev. Lett. 64 (1990) 705.
[92] H. Liu and A. A. Tseytlin, “Statistical mechanics of D0-branes and black hole
thermodynamics,” JHEP 01 (1998) 010, arXiv:hep-th/9712063.
[94] T. Banks, W. Fischler, I. R. Klebanov, and L. Susskind, “Schwarzschild black holes from
matrix theory,” Phys. Rev. Lett. 80 (1998) 226–229, arXiv:hep-th/9709091.
[95] J. W. Goodison and D. J. Toms, “The Canonical partition function for quons,”
Phys. Lett. A 195 (1994) 38, arXiv:hep-th/9410096.
[96] A. Einstein, Investigations on the Theory of the Brownian Movement. Dover Books on
Physics Series. Dover Publications, 1956.
[98] M. von Smoluchowski, “Zur kinetischen Theorie der Brownschen Molekularbewegung und
der Suspensionen,” Annalen der Physik 326 no. 14, (1906) 756–780.
[99] P. Langevin, “Sur la théorie du mouvement brownien,” C. R. Acad. Sci. (Paris) 146 (1908)
530–533.
[101] P. Berglund, T. Hübsch, and D. Minic, “Exponential hierarchy from spacetime variable
string vacua,” J. High Energ. Phys. 09 (2000) 015, arXiv:hep-th/0005162.
[103] P. Berglund, T. Hübsch, and D. Minic, “String Theory, the Dark Sector and the Hierarchy
Problem,” LHEP 2021 (2021) 186, arXiv:2010.15610 [hep-th].
[104] P. Berglund, T. Hübsch, and D. Minic, “Stringy Bubbles Solve de Sitter Troubles,”
Universe 7 no. 10, (2021) 363, arXiv:2109.01122 [hep-th].
56
[105] C. D. Froggatt and H. B. Nielsen, “Standard model criticality prediction: Top mass 173 +-
5-GeV and Higgs mass 135 +- 9-GeV,” Phys. Lett. B 368 (1996) 96–102,
arXiv:hep-ph/9511371.
[107] H. Chan, S. Tsou, and S. T. Tsou, “The Framed Standard Model (II) — A First Test
against Experiment,” Int. J. Mod. Phys. A 30 no. 30, (2015) 1530060,
arXiv:1508.04273 [hep-ph].
[109] J. F. Donoghue, K. Dutta, and A. Ross, “Quark and lepton masses and mixing in the
landscape,” Phys. Rev. D 73 (2006) 113002, arXiv:hep-ph/0511219.
[110] J. Khoury and S. S. C. Wong, “Bayesian Reasoning in Eternal Inflation: A Solution to the
Measure Problem,” arXiv:2205.11524 [hep-th].
[113] U. Aydemir, “A scale at 10 MeV, gravitational topological vacuum, and large extra
dimensions,” Universe 4 no. 7, (2018) 80, arXiv:1704.06663 [hep-ph].
[114] J.-A. Argyriadis, Y.-H. He, V. Jejjala, and D. Minic, “Dynamics of genetic code evolution:
The emergence of universality,” arXiv:1909.10405 [q-bio.OT].
[118] A. Addazi et al., “Quantum gravity phenomenology at the dawn of the multi-messenger
era—A review,” Prog. Part. Nucl. Phys. 125 (2022) 103948, arXiv:2111.05659 [hep-ph].
[120] J. Polchinski, “The Cosmological Constant and the String Landscape,” in 23rd Solvay
Conference in Physics: The Quantum Structure of Space and Time. 3, 2006.
arXiv:hep-th/0603249.
[121] N. B. Agmon, A. Bedroya, M. J. Kang, and C. Vafa, “Lectures on the string landscape and
the Swampland,” arXiv:2212.06187 [hep-th].
57
[123] Q. Wang, N. Oshita, and N. Afshordi, “Echoes from Quantum Black Holes,”
Phys. Rev. D 101 no. 2, (2020) 024031, arXiv:1905.00446 [gr-qc].
[124] V. Cardoso, V. F. Foit, and M. Kleban, “Gravitational wave echoes from black hole area
quantization,” JCAP 08 (2019) 006, arXiv:1902.10164 [hep-th].
[125] A. Strominger, “Third quantization,” Phil. Trans. Roy. Soc. London A329 (1989) 395.
https://www.jstor.org/stable/38273.
58