Let us construct a quantum version of Einstein field equation out of the classical General Relativity [
1]. Let
be a Hilbert space. Let
be a manifold, where
is an
n-dimensional differentiable manifold and
is a metric, which is either as a positive-definite section of the bundle of symmetric (covariant) 2-tensors
or as positive-definite bilinear maps,
for all
. Here,
is the subspace of
generated by elements of the form
. Let
be local coordinates in a neighborhood
U of some point of
M. In
U the vector fields
form a local basis for
and the 1-forms
form a dual basis for
, that is,
. The metric may then be written in local coordinates as
. Let
denote the Levi-Civita connection of the metric
. The Christoffel symbols are the components of the Levi-Civita connection and are defined in
U by
, and for
, we see that they are given by,
Let the curvature
-tensor
is defined by,
. Thus, the curvature tensor,
, is purely Quantum Mechanical due to (
1). Let the tensor
is the trace of
curvature tensor:
, defined by
, and the scalar curvature
is the trace of
tensor:
where
is a unit vector spanning
. Then, the Einstein-like purely quantum tensor
directly acts on a quantum space. Thus, Einstein-like field equation
, is now “purely” Quantum Mechanical for (
1). But the Ricci tensor
, thus, Einstein field equation (in quantum spacetime) should become as,
where
is Einsteinian and non-renormalizable. In the first line of (
2), all mass dimensions vanish due to
and if divergences are to be present, they could now be disposed of by the technique of renormalization (though, this will not play a role in our present discussion). Hence, (
2) should be used as a renormalizable Einstein field equation (in quantum spacetime) for general purposes. But (
2) is not sufficient to predict the properties of particles (baryonic and non-baryonic) to a great extent, such as, it never helps us to represent any kind of particle interactions, etc. So, we need to develop General Quantum Gravity (GQG), which is a formalization of quantized gravity that emerges from General Relativity through Quantum Mechanics. In the basic formalisms of GQG, we are going to develop two different aspects of GQG, such as:
In the first case, GQG gives us an Einstein field equation in a Semi-Quantum Minkowski Spacetime with Semi-Quantum Lorentzian metric tensor, whereas, the second one yields a purely Quantum Mechanical Einstein field equation in a Quantum Non-Minkowski Spacetime with Non-Lorentzian metric tensor. But in both cases, we always get the Schrödinger equation as a byproduct, though, unlike its classical form, this Schrödinger equation is now in a quantum spacetime. Additionally, this second case formalism helps us to explain Superstring/M-theory from a completely different geometric perspective.
2.1. GQG in Semi-Quantum Minkowski Spacetime
Let the line element of Minkowski spacetime,
where
is the ‘Four-momentum’, hence
for
, and
, whereas
is the ‘four-velocity’. Be careful that it is
, because
takes the ‘−’ sign. Also note that
in (
3) for the rest mass
. Thus, (
3) gives us an energy-momentum invariant line element as,
when
is the kinetic energy [
2], that is, we have a new line element as,
then, the rearrangement of (
4) gives the following equation by using (
3) as,
Let us consider the representation of a wave field
by superposition of a free particle (de Broglie wave) for (
6) as follows,
where
and
, as
, in addition, note it that we have taken here
E as total energy for (
6). Thus, from (
7), we can get the (total) energy operator
(it is analogous with, but not exactly the same as, the classical Quantum Mechanics, as it is now the total energy and related to
instead of
due to the presence of
in (
7) in either ways), the three momentum operator
, the ‘Four-momentum’ operator,
and the mass operator,
where,
is evidently relativistic, but
is no doubt Quantum Mechanical, so as,
For constant velocity, we can develop an uncertainty principle describing the intrinsic indeterminacy with which
m and
s can be determined as,
The mass-energy relation, i.e.,
, of (
4) yields the Quantum Mechanical definition for the mass operator
of (
9) for (
8) and (
10) as,
since
. Thus, (
11) tells us that the total energy of a system directly relates to its geometry, more precisely, to
. Additionally, (
11) yields the following for (
10) and
,
For the second term of
in (
9), we can rewrite (
11) by squaring its both sides after considering
as follows,
which clarifies more precisely that
is evidently relativistic, but
is definitely Quantum Mechanical. Thus, (
13), so as (
12), are very peculiar equations where the
lhs spacetime is classical relativistic but the
rhs spacetime is Quantum Mechanical, and the total energy is directly related to the
rhs spacetime. If the
lhs spacetime of (
13) changes (not more than
and not less than
, unless it is a vacuum state) and the total energy remains fixed (so as time), then the spacetime of
rhs should not remain as same as before, but changes inversely against the
lhs spacetime. Though, the increment of
rhs spacetime should not be observable, i.e., all extra dimensions would have to remain hidden inside the overall observable system, in other words, inside the
lhs observable spacetime of (
13). We will discuss it below in more details very soon.
For (
10), we can rewrite the mass-energy relation (
13) as follows,
Thus, it is somehow a kind of Klein-Gordon-like (but not an exactly similar) equation of the relativistic waves due to the above quantum scenario derived from (
7). (By the way, we will see at the very end of this Subsection that Klein-Gordon equation is a subset of a Second Order Equation of GQG in Semi-Quantum Minkowski Spacetime). This equation may yield the first order equation as,
where,
are Dirac’s gamma matrices.
Remark 1. Definitely, the above quantum scenario derived from (7) is analogous to, but not exactly similar to, the classical Quantum Mechanics since E is taken as total energy (and m is not rest mass) in (6). So, an expectation of the exactness between classical Quantum Mechanics and the present quantum scenario must lead a confusion and may yield wrong or faulty conclusions in a great extent. Readers are requested to be careful about it.
The wave field
in (
7) must satisfy the eigenfunctions for a discrete Lorentz transformation as,
when
is the complex conjugate of
. Then, using summation convention and (
10), we can write the joint state of both spacetimes as,
But, (
15) also intends to,
for
and
, where
implies
’s dependency on
and
’s dependency on
, respectively. Here,
and
have been chosen arbitrarily. Hence, the complex conjugate of (
16) is,
But, if we take,
, where
is an operator, we can say that,
, so, we may assume without any objection that,
, for some value of
. Thus,
thus, from (
16), if
, we can write as,
this yields,
if
, which gives the equivalency of (
16) and (
18). No more combinations are possible from (
16) apart from
,
and (
15) itself. The arrangement of
and
implies that
. Thus, (
16) should be rewritten by using (
15), (
16) and additionally replacing
with
as follows,
Note that,
exclusively has to depend upon spacetime. Comparing (
16) with (
18), let us say that,
where
, suppose. Since
as long as
, then it must be as follows, if
,
It is impossible to decompose (
19) and (
21) since both of their
lhs are only depended upon
∂, thus,
The spacetimes of and (so as ) are not easily dissociative even upto a very high energy scale.
Since is independent of , the spacetime of (so as ) must be an internal hidden property of the overall system (in other words, inside the observable spacetime of ). The observable spacetime is always ∂-dependent.
But, the
rhs of (
21) gives us,
where, the swap operator
for some phase
, whereas
V is a vector space. Then the corresponding eigenspaces are called the symmetric and antisymmetric subspaces and are denoted by the state spaces
and
, respectively. Note that, we have not intended here that
and
individually are two indistinguishable particles for the state spaces
and
; the above equations are just the generalization forms of their kinds, because
and
do not have distinguished (opposite) spins until otherwise they are dissociated as free particles; so, the observable spin is always the spin of
, since
is an internal hidden property of the overall system and the observable spacetime is always
∂-dependent. Thus, (
21) tells us that, if we allow
to be dissociated as a free particle at very high energy, the internal hidden spacetime of
then must be transformed into a fermionic particle, whereas, the overall
∂-dependent system remains bosonic, since, the observable spacetime is always
∂-dependent.
Similarly, (
17) yields,
this tells us that the internal hidden spacetime of
must be now bosonic, whereas, the overall system is fermionic, since, the observable spacetime is always
∂-dependent. So, whatever (
21) and (
22) want to tell us is that the
∂-dependent overall system has Supersymmetry and since the spacetimes of
and its supersymmetric partner
are not easily dissociative even upto a very high energy scale; thus,
must require extremely high energy to dissociate itself from the overall system as a free particle. Instead of being a free supersymmetric partner,
actually works quite differently inside of the observable spacetime
, though, at the same time,
is still satisfying all the properties of Supersymmetry. We will show you
’s actual purpose very soon in the below. But Supersymmetry needs extra dimensions and we should have to discuss it now.
Before proceeding with anything, we can develop a
-dependent scenario as follows,
and the complex conjugate of
is,
It is not important which state spaces satisfy such bosonic or fermionic representations of (
23) and (
24), here, the most important thing is that the overall system as a free observable particle is must not be baryonic because now only the internal hidden spacetime of
has ‘proper’ spacetime arrangement for its
∂-dependency, whereas, the overall (observable) system’s spacetime arrangement is quite ‘improper’ as it is
-dependent. Despite of
’s
∂-dependency, here, being a supersymmetric partner, if it is allowed to be free at very high energy, it must not be baryonic either. We should not be confused with it. We will discuss about its property in
Section 4 below.
The internal hidden spacetime of
in (
21) and (
22) also provides us some additional geometry for its
structures. Suppose, for
, we have,
, where, neither spacetime arrangements are matched with one another in either combinations within the curvey brackets. These ‘wrong’ arrangements must have a noticeable effect on the acceptable spacetime, i.e., its temporal part must influence over the spatially depended
, or its spatial part must influence over the temporally depended
, or vice versa. In other words, the acceptable spacetime should not have to be four-dimensional in this case. Let us check it. Suppose, for
, we can consider a dimension function (see [
3]; we follow Nagata’s work throughout this paragraph and all the proofs for this paragraph are easily obtainable by using his book [
3]),
and the space
satisfies a normal
-space. Let
be a collection in an initially
topological spacetime
, i.e.,
, which is actually hidden inside an observable
spacetime, i.e.,
(be careful here about the subscripts, do not confuse observable spacetime with hidden spacetime), and
p a point of
, then the order of
at
p should be denoted by,
, where,
is the number of members of
which contain
p. If for any finite open covering
of the topological spacetime
there exists an open covering
such that
,
, then
has covering dimension
, i.e.,
. If
can be decomposed as
for locally finite (star-finite, discrete, etc.) collections
, then
is called a
-locally finite (
-star-finite,
-discrete, etc.) collection. The topological spacetime
has strong inductive dimension
, i.e.,
, if
. If for any disjoint closed sets
F and
G of the topological spacetime
there exists an open set
U such that
, and
, where
denotes the boundary of
U, then
has strong inductive dimension
, i.e.,
. Let
V is an open set and
is a closed set of
. If
, then there exists a
-locally finite open basis
of
such that,
for every
. If a spacetime
has a
-locally finite open basis
such that,
for every
, then
. Again,
if and only if there exists a
-locally finite open basis
such that
for every
. For every subset
, for any integer
, of a spacetime
, we have,
. Hence, if and only if
for some
subsets
with
,
. For the spacetime
, we have then
. Let
A be a subset of a spacetime
and
I the unit segment. If
U is an open set of the topological product
such that
, then there exists an open set
V of
such that
, and
. Let
F be a closed set of
with
. Let
and
,
, be closed and open sets, respectively such that
, and
is locally finite. Then there exist open sets
satisfying
, and
,
, where,
, and
. Let
F,
and
satisfy the same condition as above, then there exist open sets
,
,
satisfying,
, and
for every
. Let
,
, be closed sets with
of the spacetime
. Let
be a closed collection and
a locally finite open collection such that
. Then there exists an open collection,
, such that,
, and
for every
. A mapping
f of the spacetime
into a spacetime
S is a closed (open) mapping if the image of every closed (open) set of
is closed (open) in
S. Then the continuous mappings which lower dimensions of the spacetime
should be defined as follows,
Theorem 1.
Let f be a closed continuous mapping of the spacetime onto the spacetime S such that for every . Then,
where for the space K, when , since i should not be zero in (25).
Proof. Using Theorem III.6 of [
3], we can easily prove this theorem. □
Since, the temporal axis is unaltered in Lorentz transformation, as we have already seen it in (
14), we can express the maximal continuous mapping of the
spacetime
onto the spacetime
S of (
26) as,
since
i should not be zero in (
25), if the considered state is not vacuum; then the spacetime
S definitely intends the basic structure of a 2-dimensional worldsheet
with the joint states,
for the spacetime
, where
S is a
spacetime, but for the space
K, we will like to discuss it below in more details in the Theorem 2. Obviously, a string can sweep out the 2-dimensional worldsheet
for the spacetime
.
If the internal hidden spacetime of
is considered as the
lhs spacetime of (
13) and let it to be changed from
to
when its total energy remains fixed (so as its time), then the spacetime
of
rhs of (
13) changes inversely against the spacetime of
. Since the spacetime of
is hidden inside the overall system of (
21), i.e., in other words, inside the observable spacetime of
, then the increment of
rhs spacetime
of (
13) should not be observable by any means, i.e., the extra dimensions of
remain hidden forever inside the observable spacetime of
. As these internal hidden extra dimensions inside the observable spacetime of
are considered as the representation of the spacetime
S and the space
K, thus, we can conclude,
Strings (i.e., the spacetime S for the hidden spacetime ) are natural and universal but forever hidden inside every observable system, i.e., the spacetime , in Quantum Mechanics.
Every
observable system in Quantum Mechanics must contain forever hidden extra dimensions (i.e., the space
K for the hidden spacetime
) independent of any external observer whether she/he considers any string in this system or not (for more details, see (
27) below and its following text therein).
But the space K should raise more extra hidden dimensions by a closed continuous mapping beyond by adopting the following,
Theorem 2. Let f be a closed continuous mapping of a space R onto a space K such that for each point q of K, contains at most points ; then , when and , where .
Proof. Using Theorem III.7 of [
3], we can easily prove this theorem. □
Then, we can say for the overall spacetime
that,
for which,
Note here that stringy spacetime
S vanishes in the overall spacetime
of (
27) for the space
K leaving behind the forever hidden extra dimensions
m in
. Thus, in other words, strings are experimentally unobservable forever, whereas, their actions should be mandatory in the purpose of string interactions. Also notice that Supersymmetry (now having extra dimensions
m for
due to (
28)) remains unchanged in
of (
27). Thus, with these extra dimensions, the above scenario is now perfect for Supersymmetry and String Theory without any further objections and/or adjustments.
Along with Theorem 1, what (
28) actually wants to say us is,
when
, which yields,
Since
in (
27), let the
lhs of (
29) gives,
The most disturbing thing here is that the temporal axis is a part of
S spacetime but not the part of
K space, but both
and
spaces are influenced by the (mutual) temporal axis, despite the fact that neither of them have contained any temporal axis within themselves. On the other hand, it is evidence that only an influence should not be sufficient to emerge a temporal axis within
M (or
K) space. Moreover, Theorem 2 yields no temporal axis for
M (or
K) space either. But the influenced of the temporal axis should not ease to be avoided in (
30).
From Theorem 2, if we think that the dimension of
M space depends only on
, then we should be mistaken,
M is not independent from either elements of the set
. Thinking otherwise, let
are related to new quantities
and
, differently, which are the curvilinear coordinates of
. Let the corresponding members
are determining
, then if each pair of members from the either sides of these curvilinear coordinates joining the pairs of points
and
(
) meet in points
separately, then the three points of intersection
of the pairs of coordinates
and
(
) lie on a line. Let each of the pairs of coordinates
,
(
) consists of two distinct coordinates and in which
. Let the coordinate vectors of
be denoted by
, that of
by
(
) and that of
by
(
). Then
can be represented by a linear combination of the
and
for each
, say,
Hence,
Let us choose two set of coordinates,
for
, such that
,
and
is a basis of
, whereas
, where
is the interior of
and
O is the origin, i.e.,
is admissible for
. Let the quadratic form,
say, is reduced. The last fact means that
, so that
. Since
is admissible for
, the coordinates
(
m an integer) do not belong to
. Thus,
this implies that,
Note it here that
if
and
if
do not hold due to (
31). So as,
and we can easily find that
. Here,
Similarly,
In the same way,
In the last line we have used subscripts
, which are quite different from the subscripts
we have been using yet and their purposes are quite obvious here. Since, the temporal axis is a part of
S spacetime but not the part of
K space, so both
and
spaces, as well as
and
spaces of (
31), are influenced by the (mutual) temporal axis though, neither of them have contained any temporal axis within themselves, then we can say that all axes of
and
(for
) in
K space are interrelated with the (mutual) temporal axis coming from string spacetime
S, since the temporal axis is a part of
but not the part of
K space, thus,
and
(for
) in
K space have individual existences as independent axes
and
(for
) influenced by the (mutual) temporal axis
. Let us assume that
and
(for
) in
K space have maximal weight as 1 of each dimension as an independent axis for
and
, which yields,
Hence, they have the “proper’’ dimensions. Comparing the last line of (
34) with (
32) and (
33), we can determine that if (
32) and (
33) give us some “proper’’ dimensions, then (
34) definitely gives us an “improper’’ dimension, as both
and
are depended on
and
axes, simultaneously. Since
a and
b are satisfying (
31), then
and
(for
,
) must give us “improper’’ dimensions, too. If we consider these “improper’’ dimensions
and
(for
,
) in
K space have individual existences as independent axes
and
(since they are depended on
and
axes, simultaneously) influenced by the (mutual) temporal axis
, then, on the contrary of (
35), let us assume that they have maximal weight as
of each dimension for
and
, so as they can give
, thus, we can say that,
Hence, altogether they have,
Since
, the
K space yields,
i.e.,
Thus,
has the spacetime axes as (using summation convention),
where
, whereas, other maps are obvious, for
,
and
. So, (
36) achieves,
i.e., string has eleven-dimensions by nature, that is why eleven is the maxinium spacetime dimension in which one can formulate a consistent supersymmetric theory.
Now, returning to our main purpose and using the first two terms of (
7), we can generate the following wave equation for (
6) as,
where
for
, while the ‘Four-momentum’ operator is
, and the three momentum operator is
.
Remark 2. The signature of the metric , i.e., , has been absorbed and retained unaltered by the last term of (37), as long as it satisfies (5) and (6). Thus, readers are requested to be careful not to presume space and time separately in (37), what we usually accept in the conventional Quantum Mechanics.
Again rearranging (
37) by using (
8), we may get,
or, simply discarding
, we can have the First Variance of the First Order Equation of GQG in Semi-Quantum Minkowski Spacetime as follows for
,
Evidently, (
38) may take the form
for the energy operator
. Setting the Hamiltonian operator as
, where the three momentum operator
, we can therefore have,
. Interested readers can easily check it that (
38) is nothing but the gravitational form of the classical Schrödinger equation, where
E is total energy, and now the equation has been rewritten with
along with the signature of the metric
.
It is also possible to develop a Second Variance of the First Order Equation of GQG in Semi-Quantum Minkowski Spacetime from (
37) as follows,
where
.
Now, let us multiply both sides of (
37) by
, so as,
which has the form of a general inhomogeneous Lorentz transformation (or Poincaré transformation).
Note it that (
39) is exactly equivalent to
of (
3), i.e.,
, for the ‘Four-momentum’ operator
. In other words, for (
39), we can say that the quantum line element is,
hence, by considering
, we have,
Proposition 1. From above discussion, we can deduce that:
So, we have a sufficient reason to replace the relativistic one with the quantum mechanical relation, and vice versa, from (12) as,
We will use Proposition 1 throughout our work. This proposition is quite straightforward than some commonly used textbook procedures, for example, [
4].
Remark 3.
For the Proposition 1, the relation between and of (40) should become as,
Let us consider
, etc., and let us also consider that for Proposition 1, let
would transform classical-to-quantum as,
where
is a ‘semi-quantum Lorentzian metric tensor’ in a semi-Quantum Minkowski spacetime, i.e.,
so as the ‘semi-quantum Lorentzian metric tensor’
and the classical Lorentzian metric tensor
should differ from each other and can establish a relation (
44). Without any loss of generality, we may assume that the ‘quantum metric tensor’ is symmetric:
, and
. It has an inverse matrix
whose components are themselves the components of matrix
, as their product gives:
, i.e., in terms of components,
, where,
is the Kronecker delta.
Hence, (
40) should be rewritten as,
then, let us vary the length of a curve [
5,
6,
7,
8] as,
This gives,
then the Christoffel symbol
should be defined by,
such that the Christoffel symbols are symmetric in the lower indices:
.
In this way, if we proceed further, we can develop a non renormalizable Einstein-like field equation, which is useless for our goal. To avoid this problem, let us redevelop the whole scenario from the quantum mechanical perspective as follows. Let the Christoffel symbol
should be redefined by,
such that the Christoffel symbols are symmetric in the lower indices:
. After a little exercise, we can yield the curvature tensor,
thus we find,
which satisfies the properties like symmetry, antisymmetry and cyclicity as usual, but it is not equivalent to Riemannian tensor
, i.e.,
for
in (
45). Though, the Riemann tensor (in quantum spacetime) should be derivable from (
45) as,
Now, let us take the classical-to-quantum energy momentum tensor, for example, as follows for (
42),
or the electrodynamic,
etc., and so on, hence, we have the quantum energy momentum tensors
,
, etc., those are what the graviton field couples to. Let
G is the gravitational coupling and now let us develop an unusual classical-to-quantum
G (namely
) in Planck scale using [
9] by accepting
as,
where
. Without much ado, we can easily obtain the Einstein field equation of the GQG in Semi-Quantum Minkowski Spacetime, which can give us the classical Einstein field equation while using (
46) and (
42) as follows,
Both
lhs and
rhs of (
48) independently have their mass dimensions. For the
rhs factor
of (
48), both the gravitational coupling
G (which has the dimension of inverse second order of mass) and
have lost their mass dimensions due to
. Consequently, if divergences are to be present, they could now be disposed of by the technique of renormalization. So, the last line of (
48) is not renormalizable by nature, whereas, the second equation shows us that classical Einstein field equation with
is renormalizable in quantum spacetime. (We will develop another renormalizable scenario by using a purely quantum form of gravity, which will be discussed in
Section 2.2 below.)
Remark 4. It is necessary to remember that, we should not introduce the cosmological constant Λ in (48), because we can get Dark Energy from GQG in Quantum Non-Minkowski Spacetime quite naturally (see the last equation of (73) in Section 2.2 below for more details). Otherwise, introduction of the cosmological constant Λ in (48) should intend to double entry of Dark Energy in the same gravitational field of GQG, which should obviously be faulty. Though, in (75) below, we will develop an Einstein-like field equation containing Λ, which is slightly different from the classical Einstein field equation.
As (
48) yields the non-renormalizable relation,
then, by considering d’Alembertian operator
, as well as
, we can get the Second Order Equation of GQG in Semi-Quantum Minkowski Spacetime from (
49) as,
The wavefunction
in (
50) is emphatically defining a bosonic field. Thus, as
, we can immediately develop a fermionic field (or the Third Variance of the First Order Equation of Semi-Quantum Minkowski GQG) out of (
50) for (
43) and (
12) as,
where,
are Dirac’s gamma matrices.
The classical Dirac’s equation should be derivable from (
51), but here,
is not intended to have a factor of rest (fermionic) mass, since
in (
3). Thus, we can say that Dirac’s equation is a subset of the Third Variance of the First Order Equation of GQG in Semi-Quantum Minkowski Spacetime, i.e., (
51). Similarly, we can also say that the Klein-Gordon equation is a subset of the Second Order Equation of GQG in Semi-Quantum Minkowski Spacetime, i.e., (
50). (An analogous formalism is equally applicable for the following
Section 2.2.)
2.2. GQG in Quantum Non-Minkowski Spacetime
Let the line element of Minkowski spacetime,
for the rest mass
, when,
then, rearrangement of (
52) gives,
Then, considering the representation of a wave field
by superposition of a free particle (de Broglie wave) for (
54) as follows,
we can generate the following wave equation using (
54) combining with (
52) as,
which may give,
or, simply discarding
, we can get the First Variance of the Second Order Equation of GQG in Quantum Non-Minkowski Spacetime as,
Here,
is definitely a bosonic field. But the uppermost equation of (
57) may give us the gravitational form of the classical Schrödinger equation by using (
11) for
as follows,
Putting differently,
Hence, we get the gravitational form of the classical Schrödinger equation for the total energy
E, and now it is in a
quantum spacetime.
Let us now prescribe classical-to-quantum
as follows by using Proposition 1 as,
To avoid any confusion between the classical Lorentzian metric tensor
and the quantum Lorentzian tensor of (
59), let us assume that,
This approach is quantizing gravity. The ‘quantum metric tensor’ is symmetric, i.e., , and . Components of its inverse matrix are themselves the components of matrix , namely, , where, is the Kronecker delta.
Now, applying the representation of wave field
either of (
55), or (
7), into (
54), we can get,
Note that, (
61) should be used as an alternate of (
58).
After using the first (or, second) term of mass operator
from (
9), the (
61) yields,
Interested readers may compare (
62) with (
13) and can easily retain that, in (
13),
as
and
.
Now, the Second Variance of the Second Order Equation of GQG in Quantum Non-Minkowski Spacetime from (
56) should be,
where
.
Let us try to develop an Einstein field equation, which is “purely” Quantum Mechanical (i.e., it has neither a Minkowski spacetime and not its metric is Lorentzian) in comparison to (
48). We can start with (
61), which immediately tells us that (
53) is possible to be written as,
then using (
10) into it, we have,
This line element has neither a Minkowski spacetime and not its metric is Lorentzian, since
is satisfying (
59). Note here that, the
lhs of (
64), i.e.,
, is relativistic, whereas, the
rhs components of (
64), i.e.,
, etc., are Quantum Mechanical. We can compare this line element with (
13), but this time, the relativity-to-quantum relation (or vice versa) in (
64) has been explained more uniquely and explicitly than (
13). Though, despite this advantage, we should like to achieve our goal by not accepting (
64) but by accepting the way similar to what we have already discussed in the previous subsection. Interested readers can check the workability of (
64) by themselves.
Multiplying both sides of (
56) by
and comparing it with (
52), we have the quantum line element for the ‘Four-momentum’ operator
as follows,
Similar to the
Section 2.1, after a little exercise, we can develop,
Let us develop another unusual classical-to-quantum gravitational coupling
G (namely
) in Planck scale using Proposition 1 and by accepting
as,
for
. Now, let us take the classical-to-quantum energy momentum tensor as, for example,
etc. Then, we can obtain the Einstein field equation of the GQG in Quantum Non-Minkowski Spacetime, which can give us the classical Einstein field equation as,
The first line of (
67) is renormalizable, but its last line is not renormalizable by nature. Though, (
67) yields the gravitational coupling-free field equation as,
This equation does not have any mass dimensions, so, it should be renormalizable by nature.
Interested readers can easily check that (
48) and (
67) are exactly the same expression but comprised with different components, that is, (
48) has a mixed expression of classical and quantum geometries due to
, whereas, (
67) has a purely quantum geometric expression due to
. In other words, we can say that (
48) is in a Semi-Quantum Minkowski Spacetime with Semi-Quantum Lorentzian metric tensor, whereas, (
67) is in a Quantum Non-Minkowski Spacetime with Non-Lorentzian metric tensor.
If we transform our spacetime into Planck scale, i.e.,
and
, and consider
, where
m is the mass of a certain particle and
N is a very large constant number since Planck mass is a very big number, thus,
is not considered here as the mass of a particular particle but the amount of
N number of certain particle with mass
m, then, by considering
and d’Alembertian operator
, we can get the Third Variance of the Second Order Equation of GQG in Quantum Non-Minkowski Spacetime from (
68) as follows,
Thus, the First Order Equation of GQG in Quantum Non-Minkowski Spacetime should be as follows for the fermionic mass
,
for
and
, where,
are Dirac’s gamma matrices. Definitely, (
70) gives us Dirac-like equation.
By the way, we can also rewrite (
70) as follows for the fermionic mass
,
Whatever matter satisfies (
71), it must result almost
times high population density than the critical density of matter which satisfies (
70), i.e., the fermions of (
71) have almost five-times higher critical density than the fermions of (
70), which is quite unusual. So, the fermion with mass
in (
71) is not as same as the fermion with mass
in (
70), i.e., they must be completely different particles. The only possible candidate having such characteristics as (
71) is Dark Matter, which accounts for
of the critical density in the Universe against
of the critical density of baryonic matters, in other words, the critical density of Dark Matter is almost
times higher than the critical density of baryonic matters – it exactly matches with (
71).
Again, returning to (
65) and using
of (
52) so as
for the rest mass
, we can get,
Then, considering
and d’Alembertian operator
, we have,
Thus, we can immediately develop a fermionic field equation as,
for fermionic rest mass
.
Now, let us replace
of (
52) with the Planck mass
for a certain particle with mass
, where
is a very large constant number and
as
; then
satisfies as:
, where the cosmological constant
, so as we have
for the Planck rest mass
and
, thus, for,
where,
, and following the argument cited above,
by replacing
with
for the cosmological constant
. So, by using
and by switching
right to left in the term:
, we can develop a fermionic field equation as follows,
for fermionic rest mass
as
. The interesting thing in (
74) is that Dark Energy has a direct relationship with gravity. In other words, Dark Energy would be obtainable from the breaking of particle symmetry where gravity counts, or vice versa (see,
Section 3.1 below, for example).
Note that the last equation of (
73) is definitely applicable simultaneously whether the matter is baryonic or non-baryonic.
Since (
67) and (
48) are exactly the same, by combining the last equation of (
73) with (
69), we can get a non-renormalizable field equation as.
which should be renormalizable by reintroducing
to the both terms of (
75). This field equation, which is actually a Klein-Gordon-type equation, is slightly different form the classical Einstein field equation.
Again, either from the first equation of (
73), or by placing
in (
74), we can find,
which is the Planck scale counterpart of (
71) and (
72), in other words, (
71), (
72) and (
76) are counterbalancing each other’s actions upon the Universe.
Since, the cosmological constant
, then again replacing
of (
52) with
gives us,
, for
. But, we can say,
, i.e.,
, as the
rightful and lawful ‘Dark Energy’ for relativistic
.
The grate difference between (
71) and (
76) is that the nature of the former one is non-baryonic, whereas, the later one is independent of matter’s constructive property, i.e., its effects can be observable simultaneously both in the cases of baryonic and non-baryonic matters. Another difference is that (
71) is effective at
scale, whereas, (
76) is effective at
scale, i.e., Dark Energy had originated at much earlier cosmological epochs than Dark Matter. Similarly, Dark Matter had originated at much earlier cosmological epochs than baryonic matters of (
72) at
scale. Thus, we have a quite fair chronology of the formation of cosmological matters in the Universe.
2.2.1. Superstring/M-theory
Instead of considering
of (
36), let us develop
with purely quantum spacetime axes as follows,
for
,
and
, where,
For the selection of the axis from (
77), we use fully democratic way, e.g., if
gives the
dimension, then
and
should give us the
and
dimensions, respectively, and so on.
Exclusively for this sub-subsection, we will switch to the Minkowski signature to express a point particle and a string propagating in a D-dimensional curved spacetime.
The bosonic part of the action of the
supergravity theory in 11 dimensions should be [
10],
where
. Definitely, this
supergravity is now in
spacetime of (
77) with purely quantum gravity
, which is satisfying (
59) but now with the signature
. The overall scenario is condensed inside the observable
spacetime, i.e.,
for
. In other words, the eleven-dimensional Supergravity is necessarily a natural phenomenon within the quantum spacetime of GQG.
Similarly, the bosonic part of the action of type IIA theory (
,
) should be (in the string frame),
where
,
,
and
is the dilaton. And the bosonic part of the action of type IIB theory (
,
) should be (in the string frame),
where
,
,
, while
a is the
axion and
is the dilaton. Whereas, the bosonic part of the type I action (
,
) should be,
where
is the modified field strength for the two form,
is the dilaton and
, where
A is the gauge potential in the adjoint representation of SO(32). As the two heterotic supergravity theories are obtained as the low energy limit of heterotic string theory with gauge group SO(32) and
, respectively, the bosonic part of the actions should be,
where
is the modified field strength for the two form,
is the dilaton and
, where
A is the gauge potential in the adjoint representation of SO(32) or
, respectively. In the similar way, a solitonic supergravity solution for
p-branes in 11 dimensions which interpolates between a vacuum with SO(1,2)×SO(8) symmetry yields,
where
,
, and
H is a harmonic function on the transverse space, whereas
r is the radius for the eight-dimensional space transverse to the membrane. Hence, the complete scenario of the Superstring/M-theory is condensed inside the observable
spacetime, i.e.,
for
. We know that conventionally M-theory on
does not contain any strings, however, as we have replaced Minkowski spacetime with
spacetime for (
77), hence, we have found that the new scenario of M-theory on
is now definitely contain strings, which are condensed inside the observable
spacetime, i.e.,
for
.
Let Planck mass
for a certain particle with mass
, where
N is a very large constant number. Since M-theory is the strong coupling limit of the type IIA string theory, it must be an inherently non-perturbative theory, with no arbitrary coupling constant, but only a length scale
, then the relation between this length scale and the IIA length scale and coupling can be obtained by comparing the 11 and 10-dimensional gravitational constants
and
. But we know from (
66)
in the observable
spacetime, i.e.,
for
. Then,
and
are related as,
, where,
, whereas, by accepting
,
which yields the first order equation as,
for fermionic mass
as
, when
and
are Dirac’s gamma matrices. This shows that the eleventh direction is more dynamic than what we used to think about it. Apart from this, (
80) establishes the dynamic relationship between string and the observable
spacetime’s gravitational constant
. We can again obtain the following relation using (
80),
that,
Hence, in presence of a bosonic field (
80), when the 11-dimensional radius is much smaller than the 11-dimensional length scale we effectively have a 10-dimensional theory, which is type IIA string theory. But this idea should be applicable for any dimensional radious
and any dimensional length scale since we have taken
; thus, we can rewrite (
81) as follows,
here
is a coefficient for the various energy lavels of
, for example,
. This suggests that the low energy limit of the strong coupling limit of IIA string theory (which is M-theory) must be the 11-dimensional supergravity for any dimensional radius
and any dimensional length scale in presence of a bosonic field (
80). Note it that the above all equations can yield their first order fermionic equations derivable from them, if we like so.
Let no fields depend on the
space direction
, then splitting the metric and having the three form as,
and inserting these relations into the 11-dimensional supergravity action (
78), a straightforward calculation gives the type IIA supergravity action (
79), in the string frame. If no fields depend on the
space direction
, then we have,
Hence, type IIA supergravity can be obtainable from a dimensional reduction of 11-dimensional supergravity (M-theory), where the reduced dimension should be depended on any
space direction
. So, obtaining type IIA string theory from M-theory by dimensional reduction is now not only restricted for the
space direction but universally for any
space direction.
The relations between branes in type IIA string theory and M-theory may descrbe in the following new format of tensions establishing the relation with the observable spacetime’s gravitational constant as,
Type IIA-brane:
Note that above all tension relations of type IIA-branes are highly dynamic in nature.
Since, type IIA and IIB string theories are T-dual when compactified on a small and large circle, respectively, then it must be possible to relate M-theory and type IIB string theory by using the same consideration as above. Thus, we obtain the following relations between M-theory and type IIB string theory parameters as,
This last relation can also yield its first order equation which is quite interesting in nature.
We have skipped a very interesting thing in our above discussion that would be occurred if we construct flux compactifications of M-theory to three-dimensional
spacetime preserving
supersymmetry. Here, a scalar function depending on the coordinates of the internal dimensions
(called the warp factor) is included as the explicit form for the metric ansatz as,
where,
are the coordinates of the three-dimensional spacetime
and,
are the coordinates of the internal eight-manifold
M. This opens a new scope for the understanding of string geometry.
Relating M-theory on a line interval and heterotic string theory is quite obvious now, so, we have omitted it here.