You are on page 1of 31

The Inverse Theory of Relativity, An Alternative

to Unification
Hannah Leticia Skelton

February 22, 2013 // March 1, 2024

Abstract

A new philosophical and mathematical framework is presented which


harmoniously encompasses both quantum mechanics and the general theory of
relativity.

I. INTRODUCTION

The mysterious dichotomy between quantum mechanics and the general the-
ory of relativity has been a fundamental issue in physics since the beginning of
the 20th century. In 1921, Albert Einstein was awarded the Nobel Prize in
physics for his contribution to the early formulation of quantum mechanics [10].
Ironically around that same time, he also became its greatest critic. As Einstein
was also the creator of the general theory of relativity, published in 1915, it was
his work in both theories, and his disagreement with the premise of quantum
mechanics, that put him at the historical focal point of the conflict between
them [11].

In the general theory of relativity, spacetime was shown to bend and curve
according to the mass-density of an object. This idea of the curvature of space-
time was able to explain gravitation in a whole new way. Gravity was no longer
really a force acting on an object, but an object’s ability to curve the fabric
of spacetime, thus causing other objects in its vicinity to behave according to
that curvature. One example of this is the curvature of spacetime around our
Sun which holds the planets in their orbits. The general theory of relativity was
thought to apply however, only to objects that are sufficiently large to create a
such a curvature [9].

1
In quantum mechanics, which is thought to apply only to extremely small
objects, an object’s position or momentum in spacetime cannot be known at
the same time, nor predicted with a precise certainty. All predictions in this
quantum realm become a matter of probability. Lastly, certainty in an object’s
position, momentum, or any other observable property, also has the curious
consequence of being a side effect of the observer. That is, the act of observa-
tion alters the certainty or uncertainty of the quantity being measured. This
is a sharp diversion from the classical world, including the general theory of
relativity, where objects do in fact have precise position, momentum, etc..., and
where one can make highly accurate predictions [12].

Philosophically, mathematically, and experimentally, these two theories paint


a tale of two disparate realms: general relativity relating to the very large, and
quantum mechanics relating to the very small. General relativity is the classical
theory that encompasses the curvature of spacetime; quantum mechanics is the
non-classical theory that defines physics in terms of probabilities, not certainty.

In recent times, it has been the prevailing view that these two theories must
somehow be unified because, understandably, physicists would like to have a the-
ory of physics that applies to all objects in the universe, regardless of their size.
The most prevailing attempt at unification today takes the view that quantum
mechanics is the more fundamental theory of the two. More specifically, this
prevailing view is that the universe, at its most basic, most elemental, and most
fundamental building blocks is quantum in nature; and that, therefore, quan-
tum theory can be extrapolated outwards to encompass classical mechanics, and
thus including as well the general theory of relativity.

Meanwhile, attempts at unification aside, it is still widely accepted that clas-


sical mechanics, including the general theory of relativity, has a valid place, and
validity as such, in physics. Classical mechanics still has its role in explaining
the physics of the universe at the medium/human sized level, and therefore gen-
eral relativity still explains the physics of the universe at certain large enough
scales of size.

However, the fundamental problem remains that physics at very large scales
operates according to general relativity, while at very small scales, it is quantum
in nature. Thus, it seems that these two theories, are indeed philosophically,
mathematically, and experimentally incompatible.

In this paper, I intend to flip this fundamental problem of physics on its


head. I propose that there are in fact not three theories that represent the
universe at three different scales, but that, instead, that there is one, and only
one, theory applicable at all scales, whilst we perceive that single reality from
three different perspectives.

2
II. THE NEW PROPOSAL

A. Thought experiment

The idea for this new proposal came one night whilst I was sitting outside
and looking up at the planet Jupiter with my telescope. Since my childhood, and
thanks to my father, I had always been intrigued by the dichotomy between the
general theory of relativity and quantum mechanics. But in my early twenties,
as my life headed deeper and deeper into the study of physics, I had, in effect,
become more or less obsessed with this unresolved issue.

As I looked up at Jupiter that night, I had the most peculiar thought: What
would Jupiter itself ’see’ if it could look back at me? Of course, Jupiter is not
a sentient being capable of having any thoughts at all, but the question still
seemed worthy of further consideration. I realized that, relative to Jupiter, I
was about the size of a proton. So I asked myself: Would Jupiter ’see’ me as
I would see a proton? After all, I would be so incredibly small from Jupiter’s
perspective!

Naturally, I then extended this train of thought to ask myself another ques-
tion: How would a proton see me? Again, of course, the proton is not a sentient
being capable of having thoughts of its own; but - setting that reservation aside
- as ’seen’ by the proton, I am about the size of Jupiter. So, what would the
proton ’see’ of me? Would the proton see me as I would see Jupiter? After all,
I would be so incredibly large from the proton’s perspective!

Could it be that Jupiter and the proton would, as it were, perceive me in


two such radically different ways? Could I, myself appear as, at one and the
same time or simultaneously, both the size of a proton, and the size of Jupiter
depending on the size of the observer? If so, then would it be true that all
objects in the universe are perceived in these two radically different ways, at
one and at the same time, or simultaneously, again depending upon the observer.

This was a profoundly illuminating moment for me as I realized that this


very little and accidental ’thought experiment’ might have answered my very
large, and substantive, question concerning the resolution of the conflict between
the general theory of relativity and quantum mechanics.

3
B. Theories of relativity

Thus by extension of the thought experiment, the new proposal was born. I
first determined that physics needed to be flipped on its head. That is, instead of
considering the universe in terms of small, medium, and large objects, I needed
to instead think of it in terms of small, medium, and large observers of any and
all objects in the universe. Any object, including ourselves, can in fact be seen
as being small, medium, or large, depending upon the size of its observer at
any given moment in time. Further, any object can be seen as small, medium,
and large, simultaneously, and at all times, by considering the perspectives of
all possible observers at any moment in time.

This led to the radical leap, but also the common sense step, that Jupiter
may in fact perceive me as I would perceive a proton, that is, behaving according
to quantum mechanics. After all, somewhat as an extension of Einstein’s equiv-
alence postulate [9], there was no actual way to discern a difference between an
object of Jupiter’s size looking at me, and object of my size, looking at a proton.
The size ratios were close to the same, and thus it would be impossible for an
observer to distinguish oneself as being in one or the other frames of reference.

The second step forward was to hypothesize, conversely, that the proton may
in fact perceive me as I would perceive Jupiter - in other words, according to
the general theory of relativity.

Thus I concluded that it was indeed plausible, through this inverted, observer-
based take on physics, that I, a human being, could be following quantum me-
chanics and the general theory of relativity simultaneously. And therefore I
concluded that any object could be seen as small enough relative to its observer
to be following quantum mechanics, and large enough, with respect to its ob-
server, to be following general relativity. Therefore, any and all objects in the
universe could be said to be simultaneously behaving according to quantum me-
chanics and general relativity, even though the size of the observer with respect
to the object at any moment in time would determine which theory the observer
would reflect.

Thus quantum mechanics and the general theory of relativity are no longer
only object based theories of small, medium, and large sized objects. They are
indeed this; but they are also more than this. They are theories of relativity,
based on the size of any observer with respect to any object, and they apply
simultaneously to all objects in the universe at all times.

4
C. All Possible observers

One question that should be asked here is why it would be necessary, or even
reasonable, to consider all possible observers of a given object? After all, none
of them are sentient, much less capable of understanding the laws of physics! So,
why should the perspectives of Jupiter and/or the proton even matter, let alone
be considered? The answer is, quite simply, because it is unreasonable and in
fact, untrue to assert that an object is small, medium, or large just because it
is so with respect to the size of a human being. The human being is, perhaps,
’divine’ in some respects, but our size is not one of them! Further, it is anything
but ’divine’ that our laws of physics seem to divide only too well into our human
versions of the small, medium, and large sizes of objects around us. Instead, it
is perhaps our own failure to look beyond our human bias - beyond our unique
size and place in the universe - that has caused this apparent dichotomy between
the physics of the small and the physics of the large. We cannot look for an
objective theory of physics based on a single subjective perspective, that is,
based on the the human perspective alone.

One has to see our universe from all possible perspectives and vantage points,
in order to overcome and see beyond our unique perspective. Even then, a purely
objective, or noumenal, knowledge of physics may not be possible [7] [14]. To
achieve at least a subjectively harmonious, phenomenal theory, we must consider
the views of other potential observers. Jupiter, or the proton, would look out at
the universe and see quite different perspectives. To Jupiter, we are the proton,
and to the proton, we are Jupiter!

D. Extending the general theory of relativity

If a human being, and thus any object, can be seen as both very large and
very small simultaneously, and thus behaving according to both the general the-
ory of relativity and quantum mechanics simultaneously, then we must consider
the question of how these two theories might fit together in this new framework.

My hypothesis is that general relativity and quantum mechanics are in a


sense inverses of one another. However, I also assert that it is general relativity
that is in fact the more fundamental theory of the two. I say this because if
one undertakes this quest from the viewpoint of the quantum mechanics side,
and then from viewpoint of general relativity, it becomes apparent, both mathe-
matically and philosophically, that general relativity leads to the less ridiculous
conclusions. Quantum mechanics as a starting point leads to the notion that
gravity, and thus spacetime itself, and even consciousness, may be quantized.
But, starting from the foundation of general relativity, one is not led to such

5
far-fetched conclusions; instead, light and clarity may even be shed onto some
of the mysterious - if not seemingly absurd - aspects of quantum mechanics.
Lastly, the prevailing idea that quantum mechanics should be the dominant
theory simply because it is how the smallest (relative to us) objects appear to
behave, should not trump the fundamental role that the very fabric of spacetime
itself plays in the universe and in human consciousness. The fabric of space-
time both contains and, in a sense, transcends the objects in it, as well as their
size. Thus I consider that this should be reason enough to make the fabric of
spacetime, and therefore the general theory of relativity, the more fundamental
of the two theories.

Therefore, taking general relativity as my starting point, I propose that the


curvature of spacetime is the reason for the appearance of the phenomenal world
as we see it in the general theory of relativity itself, in the everyday world of
classical mechanics, and even in the world of quantum mechanics.

Additionally, using the general theory of relativity as my starting point, I


further propose that spacetime curvature, being so fundamental, is not limited
only to sufficiently large (relative to us) bodies; rather that spacetime is in fact
curved around all bodies, regardless of their size. I suspect that it is precisely
this curvature of spacetime which creates the appearance of quantum mechanics,
and which could shed light on much of its mysterious nature.

E. Mathematical structure

We begin with Jupiter, which has spacetime curved around itself according
to the general theory of relativity. Then, rather than considering the size of a
smaller object, and thus considering a proton for example, as in the usual way
of organizing physics, one instead increases the size of the observer until Jupiter
becomes the apparent size of the proton.

We will show here that locally, the motion of Jupiter, from the perspective
of a much larger observer, is isomorphic to the motion of the proton, from our
usual human perspective. This provides a mathematical framework that can
act as a starting point to accomodate this theory.

It should be noted that we are disregarding the charge of a proton. Only


the size of the proton, as a free particle, is relevant.

6
1. Mathematical setting

We let our mathematical setting for the curvature of spacetime around


Jupiter be a tangent vector bundle π : E → M . Here M is Minkowski space-
time, that is, the 4-dimensional, orientable, Lorenztian manifold (M, g) with
Levi-Civita connection D, andS E is the union
S of all tangent vector spaces, Tp M ,
over M . More precisely, E = p∈M Ep = p∈M Tp M . Lastly, we define a sec-
tion of the tangent vector bundle as a mapping, s : M → E, such that s(p) ∈ Ep
for any p ∈ M [1].

2. Spacetime curvature and Lie algebra

In this section we want to show that the curvature of spacetime around


Jupiter, when its apparent size approaches zero, that is, as the observer of
Jupiter grows increasingly larger with respect to the planet, is an element of the
Lie algebra so(3, 1).
We begin with a point p ∈ M and then we want to parallel transport a
vector v(t) starting in Ep , around a square on the surface of Jupiter. We parallel
transport vector v(t) in the xµ -xν plane. We let v(0) = v.
However we start out by deriving parallel transport on a more general frame-
work, the vector bundle, rather than the tangent vector bundle.

(a) Holonomy and parallel transport. Let us define the holonomy on a vector
bundle. Consider a closed smooth path γ : [0, T ] → M , at point p ∈ M . Thus let
γ(0) = γ(T ) = p in a manifold M with E being a vector bundle with connection
D. Then for some u ∈ Ep , let H(γ, D)u be the result of parallel transporting u
around the closed path γ. We call H(γ, D) the holonomy of the vector bundle
[1].
Thus by definition of the holonomy, we know that our initial vector v ∈ Ep ,
upon parallel transport around a closed smooth path, is consequently acted upon
by the holonomy group element H(γ, D) such that v will come back rotated,
such that v(t) will be,

v(t) = H(γ, D)v.

As we are working with a 4-dimensional, orientable, Lorentzian manifold,


this holonomy is an element of the Lie group SO(3, 1) [2].

(b) Explaining parallel transport. The equation for parallel transport and
its mathematical setting should also be noted and more fully established within
the framework of the general vector bundle.

7
Consider a vector v starting in Ep , that is being parallel transported along
a path γ(t) in a vector bundle. As we are traversing different sections of the
bundle, each with different vector spaces, parallel transport requires keeping
the covariant derivative of the vector v(t), in the direction of γ(t), that is γ 0 (t),
equal to zero. Thus we have as our condition for parallel transport,

Dγ 0 (t) v(t) = 0.

More generally, the covariant derivative of any vector field, u, on M , is


assigned by the connection D, as a function between sections of the vector
bundle. Thus, given any section s in E, and vector field on u, on M , in a vector
bundle, we call Du s the covariant derivative of s in the direction u.

In other words, the covariant derivative, is really another way of expressing


the connection D, and it allows us to differentiate across sections of a vector
bundle.

Since the covariant derivative of our section, s, in E, is defined on an open


set, but our vector u in M , only lives on the curve γ(t), we will need a local
trivialization. Thus, we have,

E|Uγ ∼
= Uγ x V ,

of E, over a neighborhood U of γ(t).

Our connection, or covariant derivative across sections, is defined here as the


standard flat connection plus a vector potential A [1],

Dµ s = ∂µ s + Aµ s.

Here, the vector potential, Aµ , are the components of the vector potential A,
across sections. The vector potential A, and its components, Aµ , are elements
of the Lie algebra, which in this case is so(3, 1).

Thus, analogously, we define the covariant derivative on this vector bundle,


locally as,

Dγ 0 (t) v(t) = d
dt v(t) + A(γ 0 (t))v(t).

Thus, starting with our vector v ∈ Ep , we parallel transport it along γ(t) in


order to find, v(t) ∈ Eγ(t) .

We let,

v(0) = v and Dγ 0 (t) v(t) = 0.

8
Thus we just need to solve the differential equation,

Dγ 0 (t) v(t) = 0,

or more specifically, that is, to solve,

d
dt v(t) + A(γ 0 (t))v(t) = 0.

which has the solution,

A(γ 0 (s))ds
Rt
v(t) = P e− 0 v,

and whose existence and uniqueness is guaranteed by the fundamental the-


orem of ordinary differential equations.

(c) Performing parallel transport around a loop. In order to parallel trans-


port around a loop, we start by using our formula for parallel transport,

A(γ 0 (s))ds
Rt
v(t) = P e− 0 v.

We then take the path integral in the vector potential A, around path, which
we define as a square loop. We let this square have side length  and we define
the path γ by 0 ≤ xµ ≤  and 0 ≤ xν ≤ . Thus, we take the path integral,

A(γ 0 (s))ds
R1
P e− 0

whose path ordered exponential is as follows [1],


(−1)n R1
Σ∞ 0 n
n=0 n! P ( 0 A(γ (s))ds) .

Since we are working only on a very small local loop, we will only take this
path integral up to second order. Thus, we expand the path ordered exponential
as,
R1 R1
1− 0
A(γ 0 (s))ds + 12 ( 0 A(γ 0 (s))ds)2 .

We now define our loop as a more specific path, γ(t). We let,

9
if 0 ≤ t ≤ 41 ;


 (4t, 0)



 1
 (, (4t − 1)) if < t ≤ 12 ;


4
γ(t) =
1
((3 − 4t), ) if < t ≤ 34 ;




 2


 3
(0, 4(1 − t)) if < t ≤ 1.

4

Thus γ 0 (t) is,

if 0 ≤ t ≤ 41 ;


 4∂µ



 1
 4∂ν if < t ≤ 21 ;


4
0
γ (t) =
1
−4∂µ if < t ≤ 43 ;




 2


 3
−4∂ν if < t ≤ 1.

4

Our first integral yields,


R1 1 1
A(γ 0 (s))ds =
R R
0 0
4
4Aµ (4s, 0)ds + 1
2
4Aν (, (4s − 1))ds −
4

R 3 R1
1
4
4Aµ ((3 − 4s), )ds − 3 4Aν (0, 4(1 − s))ds.
2 4

We can expand these integrands as a Taylor expansion, to first order in ,

Aµ (4s, 0) = Aµ + 4s∂µ Aµ

Aν (, (4s − 1)) = Aν + ∂µ Aν + (4s − 1)∂ν Aν

Aµ ((3 − 4s), ) = Aµ + (3 − 4s)∂µ Aµ + ∂ν Aµ

Aν (0, 4(1 − s)) = Aν + 4(1 − s)∂ν Aν

which becomes, upon a Taylor expansion of the integrands,

10
R1 1
A(γ 0 (s))ds = (4Aµ + 162 ∂µ Aµ s)ds +
R 4
0 0

1
(4Aν + 42 ∂µ Aν + 42 (4s − 1)∂ν Aν )ds −
R 2
1
4

3
4Aµ + 42 (3 − 4s)∂µ Aµ + 42 ∂ν Aµ )ds −
R 4
1
2

R1
3 (4Aν + 162 (1 − s)∂ν Aν )ds,
4

which becomes after integration,

= Aµ + 12 2 ∂µ Aµ + Aν + 2 ∂µ Aν + 12 2 ∂ν Aν − Aµ

− 12 2 ∂µ Aµ − 2 ∂ν Aµ − Aν − 12 2 ∂ν Aν .

This simplifies to,

2 (∂µ Aν − ∂ν Aµ ).

Next, our second integral can be written,


R1 R 1 R s1
1
2( 0 A(γ 0 (s))ds)2 = 0 0
A(γ 0 (s1 ))A(γ 0 (s2 ))ds2 ds1 .

The inner integral yields,

R s1
0
A(γ 0 (s2 ))ds2 =

if 0 ≤ t ≤ 41 ;


 4s1 Aµ



 1
 Aµ + (4s1 − 1)Aν if < t ≤ 21 ;


4

1
Aµ + Aν − (4s1 − 2)Aµ if < t ≤ 43 ;




 2


 3
Aν − (4s1 − 3)Aν if < t ≤ 1.

4

11
Then for outer integral we have,

R1 R s1 1
ds2 A(γ 0 (s1 ))A(γ 0 (s2 )) =
R
0
ds1 0
4
0
4Aµ (4s1 Aµ )ds+
R 1
1
2
4Aν (Aµ + (4s1 − 1)Aν )ds −
4
R 3
1
4
4Aµ (Aµ + Aν − (4s1 − 2)Aµ )ds −
2
R1
3 4Aν (Aν − (4s1 − 3)Aν )ds,
4

which after integration, becomes,

= 21 2 A2µ + 2 Aν Aµ + 21 2 A2ν − 2 A2µ − 2 Aµ Aν

+ 12 2 A2µ − 2 A2ν + 12 2 A2ν .

This simplifies to,

2 [Aν , Aµ ].

Thus, letting the size of the square tend towards zero, that is, as  → 0, our
terms simplify down to,

1 − 2 (∂µ Aν − ∂ν Aµ ) − 2 [Aµ , Aν ],

which by the definition of spacetime curvature, in terms of the vector potential,A,

Fµν = ∂µ Aν − ∂ν Aµ + [Aµ , Aν ]

yields then,

1 − 2 Fµν .

As stated earlier, upon parallel transport around the square loop, our vector
v is acted upon by an element of the holonomy group, H(γ, D), which is an
element of the Lie group SO(3, 1).

12
However, as Jupiter and the square, together and in equal measure, become
infinitely small with respect to their observer, this holonomy group becomes
a measure of the curvature of spacetime and thus also becomes a differential
element of that holonomy group, thus becoming an element of the Lie algebra
so(3, 1). This is a result of the Ambrose-Singer theorem [2].

Thus,

v(t) = H(γ, D)v

becomes,

v(t) = (1 − 2 Fµν )v

where, (1 − 2 Fµν ) ∈ so(3, 1).

It is important to note here that it is the apparent size of Jupiter, as well as


the loop on it, that appear to be getting smaller. It is not simply a square on
the surface of Jupiter that is getting smaller relative to Jupiter. It is the square
loop and Jupiter, together, that appear to be getting smaller and smaller, as
the size of the observer gets larger and larger. If the square was simply getting
smaller relative to Jupiter, the curvature would become flat on an infinitely
small element of the tangent space on Jupiter’s surface. However, when the
planet itself, as well as the square on it, appear to get smaller and smaller,
together and in equal measure, the geometry enclosed by the loop is unchanged,
and thus the curvature is preserved, though the loop appears to get smaller and
smaller.

(d) The more specific case of the general theory of relativity. It should also be
mentioned here that in the general theory of relativity, we are actually dealing
with a more particular, though very much analogous situation.

Instead of our more general equation for parallel transport on a vector bun-
dle, defined with a connection D of a vector v(t), along a path γ(t) in M ,

Dγ 0 (t) v(t) = d
dt v(t) + A(γ 0 (t))v(t),

we are in general relativity, considering parallel transport as defined on the


tangent bundle, defined by the Levi-Civita connection. We use the Levi-Civita
connection because our base manifold is endowed with a metric and the con-
nection is both metric preserving and torsion free. Thus, for the Levi-Civita
connection, we are dealing with parallel transport of specifically, a tangent vec-
tor along a path γ(t), defined as [1],

13
λ
Dγ 0 (t) v µ (t) = d µ
dt v (t) + Γµνλ dγ ν
dt v (t).

That is, with v µ (t) equal to the tangent vector, γ 0 (t), being parallel trans-
ported along the path γ(t).

We see that instead of the vector potential A from the vector bundle, with
general relativity, with the tangent bundle, we are using the Christoffel symbols
Γ.

The vector potential, A, results in the taking of the covariant derivative, D


along sections of the vector bundle. In other words, the vector potential are the
coefficients of the connection Dµ ej , expressed uniquely as a linear combination
of the sections ei , with functions on U ⊆ M as the coefficients. Thus we define
the functions Aiµj on U as [1],

Dµ ej = Aiµj ei ,

where, Dµ = D∂µ ,

noting that ∂µ is the corresponding basis of coordinate vector fields, and


that ei be a basis of sections of E over U .

The Christoffel symbol too, results as a consequence of taking the covari-


ant derivative of sections of the vector bundle, but they are specifically the
coefficents of the Levi-Civita connection on the tangent bundle. They are de-
fined as the unique connection coefficients of the Levi-Civita connection. The
Levi-Civita connection on M, taken in the coordinate direction ei , are,

∇i ej = Γkij ek .

where ∇i = ∇ei ,

and noting that ei = ∂i is a local coordinate basis.

Thus parallel transport in each case, respectively, is,

Dγ 0 (t) v(t) = 0,

and,

∇γ 0 (t) γ 0 (t) = 0.

expanded more specifically as [1],

14
d
dt v(t) + A(γ 0 (t))v(t) = 0

and,
λ
d 0
dt γ (t) + Γµνλ dγ 0
dt γ (t) = 0

respectively.

Lastly, the holonomy after parallel transport, again to second order, is,

(1 − 2 Rµνδ
α
)∂α

where ∂α is the tangent vector, being parallel transported around a square


again, of size , in the xµ − xν plane,

rather than,

(1 − 2 Fµν )v,

with v being the vector in section of the vector bundle, Ep , and whose parallel
transport around xµ − xν , was explained previously [1] [9].

These situations are analogous, though in the case of the general theory of
relativity, we are using the notation that is specific to the tangent vector bundle
and the Levi-Civita connection.

We are still using the SO(3, 1) Lie group, and also still using the Ambrose-
Singer theorem. Thus our curvature still lies in an infinitesimal element of
the Lie group/holonomy group and is therefore still an element of the so(3, 1)
Lie algebra. Though we acknowledge the differences with the general theory of
relativity, we use the more general notation with A, for a connection on a vector
bundle, just so that we may more clearly illustrate our argument.

3. Holonomies along a path

In this section we let Jupiter, at this apparent small size, travel along a
maximally smooth piecewise path. We consider its curvature one infinitely
small square loop at a time, from some point p to some point q, where p, q
∈ M . This yields a path ordered product of holonomies along the path, yielding
a net holonomy for the entire path [1],

H(γ, D) = H(γn , D)...H(γ1 , D).

15
Thus the result of parallel transporting the vector in loops along a path from
p to q, gives a net holonomy along the path,

v(t) = H(γ, D)v.

where H(γ, D) is a linear map,

H(γ, D) : Ep → Eq .

Equivalently, our equation for parallel transport along a path, the result
yields,

A(γ 0 (s))ds
Rt
v(t) = P e− 0 v.
0
Rt
Thus H(γ, D) is equivalent to P e− 0 A(γ (s))ds , along the entire path, and
both are elements of the Lie group SO(3, 1), and with A(γ 0 (t)), an element of
the Lie algebra so(3, 1).

Furthermore, each square loop element, is an element of the Lie algebra


so(3, 1), along the path of loops, for the apparent size of Jupiter, traveling
along that path.

4. The local isomorphism with quantum mechanics

Lastly, we show in this section, through a local Lie algebra isomorphism,


that the equation of motion of Jupiter along a path, from the perspective of a
much larger observer, is differentially identical to the equation of motion of a
proton along a path in quantum mechanics.

(a) The general theory of relativity side. Our differential equation for par-
allel transport of Jupiter, being the apparent size of a proton, along a path γ(t)
being,

Dγ 0 (t) v(t) = 0

which becomes,

d
dt v(t) = −A(γ 0 (t))v(t)

where A(γ 0 (t)) is an element of the Lie algebra so(3, 1).

16
The solution of which is,

A(γ 0 (s))ds
Rt
v(t) = P e− 0 v

A(γ 0 (s))ds
Rt
where e− 0 is an element of Lie group SO(3, 1).

Remembering that this is also an element of holonomy group, H(γ, D), and
that as the holonomy / Lie group element shrinks down to zero, or in our case
specifically, as the apparent size of Jupiter shrinks down to zero, it becomes a
function of the spacetime curvature, and thus, also a differential element of the
holonomy/Lie group, and thus becomes an element of the Lie algebra so(3, 1),
for differential motion.

(b) The quantum mechanics side. Next, we consider our differential equation
for the motion of the proton, from quantum mechanics being,

d
i~ dt |ψ(t)i = Ĥ |ψ(t)i

where Ĥ is an element of the Lie algebra sl(2, C) [3].

And with solution,


iĤt
|ψ(t)i = e− ~ |ψ(0)i.
iĤt
Where the exponential term, e− ~ is an element of the Lie group SL(2, C).

It should be noted that we know Ĥ is in the Lie algebra sl(2, C), because
the general form of the propagator for the Hamiltonian is [3],

1 λ
H(λ, ω) = 2m (p
2
+ x2 ) + 12 mω 2 x2

and letting ω = 0, we get simply,

1 2 λ
H(λ, ω) = 2m (p + x2 ).

Additionally, the following commutation relations do provide a realization


of the Lie algebra, sl(2, C),

λ
[x2 , p2 + x2 ] = 2i~(xp + px)

[x2 , xp + px] = 4i~(x2 )

λ λ
[p2 + x2 , xp + px] = −4i~(p2 + x2 ).

17
with the relation, [x, p] = i~.

Thus we have,

1 2 λ
H0 (λ) = 2m (p + x2 )

λ
where (p2 + x2 ) ∈ sl(2, C)

1 λ
and exp(−i 2m~ (p2 + x2 )t) ∈ SL(2, C).

We can then take the case for the free particle propagator, thus we let λ = 0
Our propagator for the Hamiltonian becomes just,

1 2
H0 (λ) = 2m (p )

which leads us to the well known unitary operator for the free particle in
quantum mechanics [3],

U ˆ(t) = e− ~ H0 (λ) ,
it

or simply,
itp2
U ˆ(t) = e− 2m~ ,
2
itp
and where − 2m~ is an element of the Lie algebra sl(2, C).

(c) The local isomorphism. Now, locally these two Lie groups, SO(3, 1),
and SL(2, C) are isomorphic, or rather, the Lie algebra so(3, 1) is isomorphic
to the Lie algebra sl(2, C)

We show that there is Lie algebra isomorphism, φ : sl(2, C) → so(3, 1), by


showing that the map, φ, is a bijective function as well as that the Lie bracket
is preserved.

Thus we define a mapping, φ : sl(2, C) → so(3, 1),

More specifically we first consider the basis for the Lie algebra sl(2, C) as [6],

( 21 σ1 , 21 σ2 , 12 σ3 , 2i σ1 , 2i σ2 , 2i σ3 ) = (ji , jj , jk , ki , kj , kk )

where the σ represents the Pauli matrices and with the commutation rela-
tion, [σi , σj ] = 2iijk σk .

18
Next the basis for the Lie algebra, so(3, 1), with the metric defined as η =
diag(−1, 1, 1, 1), we have,

(Ji , Jj , Jk , Ki , Kj , Kk ),

where,

   
0 0 0 0 0 1 0 0
0 0 0 0 1 0 0 0
Ji = i   Ki = i  
0 0 0 −1 0 0 0 0
0 0 1 0 0 0 0 0
   
0 0 0 0 0 0 1 0
0 0 0 1 0 0 0 0
Jj = i 
0
 Kj = i  
0 0 0 1 0 0 0
0 −1 0 0 0 0 0 0
   
0 0 0 0 0 0 0 1
0 0 −1 0 0 0 0 0
Jk = i 
0
 Kk = i  
1 0 0 0 0 0 0
0 0 0 0 1 0 0 0

Then the mappings, defined as φ( 12 σi ) → Ji , and as, φ( 2i σi ) → Ki ,

by extending linearity, are bijective.

Lastly these mappings also preserve the Lie bracket, given that the commu-
tation relations for sl(2, C) are,

[ji , jj ] = iijk jk , [ji , kj ] = iijk kk , [ki , kj ] = −iijk jk ,

and the commutation relations for so(3, 1), are [4],

[Ji , Jj ] = iijk Jk , [Ji , Kj ] = iijk Kk [Ki , Kj ] = −iijk Jk

and thus,

f ([ 12 σi , 21 σj ]) = f ( 14 [σi , σj ])

= f ( 12 iijk σk )

= iijk Jk

19
= [Ji , Jj ]

= [f ( 12 σi ), f ( 12 σj )].

and, similarly,

f ([ 2i σi , 2i σj ]) = f (− 14 [σi , σj ])

= f (− 12 iijk σk )

= −iijk Jk

= [Ki , Kj ]

= [f ( 2i σi ), f ( 2i σj )].

Thus, since the mapping between so(3, 1) and sl(2, C) is bijective and the
Lie bracket is preserved, we have a Lie algebra isomorphism. Thus, we can say
that the element A(γ 0 (t)) ∈ so(3, 1), is isomorphic to the Hamiltonian element,
it
~ H0 (λ) ∈ sl(2, C).

Further, since our two Lie algebras are isomorphic, we can say that the two
Lie groups here, SL(2, C) and SO(3, 1), are locally isomorphic. In the physical
sense, we can say that the motion of Jupiter, from the perspective of a much
larger observer, is locally isomorphic, or differentially equivalent, to the motion
of a proton, from the human perspective. Thus, given the same appropriate
differences in size between observer and object, the motion of Jupiter along a
path becomes structurally equivalent, locally, to the motion of a proton.

III. FURTHER THOUGHTS

A. Transformations on Minkowski spacetime

In addition to this isomorphism between so(3, 1) and sl(2, c), it should be


additionally noted that the action of the SL(2, C) Lie group on a 4-dimensional
Minkowski spacetime, M ,

σ(A, x) = AX(x)A† ,

represents a Lorentz transformation, where A ∈ SL(2, C), and X(x) ∈ SL(2, C).

20
This occurs when one first rewrites a point in Minkowski spacetime, x =
(x0 , x1 , x2 , x3 ), in terms of Pauli matrices, σi = (I, σ1 , σ2 , σ3 ), as,

X(x) = x0 I + x1 σ1 + x2 σ2 + x3 σ3 ,

which becomes,
 0
x + x3 x1 − ix2

X(x) = 1 .
x + ix2 x0 − x3
The Minkowski norm and determinant are preserved here, as,

detX(x) = (x0 )2 − (x1 )2 − (x2 )2 − (x3 )2 = −xµ xµ ,

and thus,



 > 0 if x is timelike;



detX(x) : = 0 if x is on the light cone;




< 0 if x is spacelike;

Hence, the action, σ(A, x), where A ∈ SL(2, C) and x ∈ M , also preserves
the Minkowski norm and the determinant as,

detσ(A, x) = detAX(x)A† = detX(x) = −xµ xµ

where detA = detA† = 1 .

Thus, we have a Lorentz transformation [2].

Furthermore and conversely, a unique vector in spacetime, x, can also be


written in terms of the 2 x 2 Hermitian matrix, as

xµ = 12 tr(σµ X).

While there is an isomorphism between elements of spacetime, in M , and


the elements of the 2 x 2 Hermitian matrices, there is only a two-to-one homo-
morphism between a SL(2, C) and SO(3, 1), as A and −A ∈ SL(2, C) give the
same element of SO(3, 1) [2].

21
B. A note on the Ambrose-Singer theorem

The Ambrose-Singer theorem: Let P (M, G) be a G-bundle over a connected


manifold M. The Lie algebra h of the holonomy group Φuo of a point u0 ∈ P
agrees with the subalgebra of g spanned by elements of the form,

Ωu (X, Y ) for X, Y ∈ Hu P ,

where u ∈ P is a point on the same horizontal lift as u0 [2].

The Ambrose-Singer theorem, though it applies to vector bundles and to the


curvature of spacetime in the general theory of relativity, also more generally
applies to principal G-bundles. While our explanation of parallel transport,
curvature, and holonomy explained previously in the context of vector, and
tangent vector bundles, is advantageous in that it is in keeping with the standard
convention and also helps to provide a mental picture of the physical situation,
we will also here discuss the broader context within principal G-bundles.

This broader context is helpful in that it shows more clearly the relationship
between the curvature and the Lie algebra. However, and perhaps most impor-
tantly, it also shows how this new framework of the general theory of relativity,
interpreted within the context of principal G-bundles, can be thought of in a
way that is much closer in structure to the Yang-Mills theory, which occurs
naturally as a principal G-bundle. The Yang-Mills theory is the underlying
framework for the Standard Model of today. It is also the framework for clas-
sical electromagnetism, whose unification with the general theory of relativity,
was the beckoning question of the past for Einstein and his colleagues.

This new relationship between the general theory of relativity and quan-
tum mechanics, shows how the general theory of relativity could be the new
philosophical foundation, but also, through this principal G-bundle structure,
the underlying mathematical foundation as well. That is, perhaps, gravity, or
rather, the curvature of spacetime is not one of the fundamental forces after all,
but is instead, the underlying philosophical and mathematical structure.

1. The general theory of relativity in principal G-bundles

We thus briefly explain our new framework within this broader setting of
the principal G-bundle.

(a) The principal G-bundle. A principal G-bundle is defined as as a fibre


bundle, π : E → M , such that p = π(u) for p ∈ M and u ∈ E. However, the

22
fibers are not simply vectors, as in a vector bundle, as discussed earlier, but are
actually identical to the structure group, or Lie group, G, of the base manifold,
M . It is also denoted by P (M, G) and often called a G-bundle over M [2.]

A fibre of the principal G-bundle is defined as Gp = {u ∈ E|π(u) = p}.

A section of a principal G-bundle is a smooth map σ : M → E such that for


any p ∈ M , σ(p) ∈ Ep .

Technically, our vector bundle from above in our theory, is a principal G-


bundle because we are using the Lie group SO(3, 1) [2].

(b) The vertical and horizontal subspaces. Again, as described earlier, we


still consider parallel transport around an infinitely small loop, as both Jupiter
and the loop on it grow infinitely small in equal measure with respect to their
observer. However, this time, since parallel transport is being considered on a
principal G-bundle, it must be thought of in slightly different terms. In this
context, the tangent space on E, Tu E, must be divided into vertical and hori-
zontal subspaces, before we can define the connection and other following and
further constructed terms.

The vertical subspace Vu E is defined as a subspace of Tu E which is tangent


to Gp at u,

More specifically, we take an element A of g, and by the right action,

Rexp(tA) u = uexp(tA)

we define a curve through u, in E.

Then since π(u) = π(uexp(tA)) = p,

we can define a vector A] ∈ Tu E by,

d
A] f (u) = dt f (uexp(tA))|t=0 ,

where f : E → R, is an arbitrary smooth function. This vector A] is


therefore tangent to Gp at u, hence A] ∈ Vu E.

It should be noted that there is a vector space isomorphism ] : g → Vu E,


given by, A → A] .

Meanwhile the horizontal subspace, Hu E, is defined as the complement of


Vu E [2].

23
(c) The connection. The connection here is defined as the Lie algebra valued
one-form, ω, such that, ω ∈ g ⊗ T ∗ E.

This connection can be thought of as a projection of Tu E onto the vertical


component of the tangent space, Vu E, and which is isomorphic to the Lie algebra
g. More specifically, there is an isomorphism, Vu P ' g, defined by ω(A] ) = A.

Thus, an element of horizontal subspace is actually a vector that has zero


projection onto the vertical subspace, or Lie algebra elements. Thus we can
more specifically define the horizontal subspace as,

Hu E = {X ∈ Tu E|ω(X) = 0}.

The connection one-form defined here, ω, splits the tangent space Tu E into
Hu E ⊕ Vu E, and is known as the Ehresmann connection [2].

(d) Parallel transport. When we parallel transport our vectors in this con-
text, it is only the vectors in the horizontal subspace that are transported. The
condition for parallel transport becomes, ω(X) = 0 for all X ∈ Hu E. Specifi-
cally, these elements of the horizontal subspace above the path γ(t), are elements
of what is called horizontal lift.

A curve in γ̃ : [0, 1] → E is said to be a horizontal lift of γ if πγ̃ = γ, and


the tangent vector to γ̃(t) always belongs to Hγ̃(t) E, that is, for some vector X̃,
tangent to γ̃, we have, X̃ = γ̃∗ X.

Thus, letting X̃ be a tangent vector to γ̃, and given our condition for parallel
transport, ω(X̃) = 0, our differential equation becomes,

dgi (t)
= −ω(σi∗ X)gi (t),
dt
or,

dgi (t)
= −Ai (X)gi (t),
dt

where Ai (X) = σi∗ ω(X) = ω(σi∗ X)

and where, σi is a local section on a Ui of M .

This equation, whose existence and uniqueness is again guaranteed by the


fundamental theorem of ordinary differential equations, has solution,

24
R γ(t)
gi (γ(t)) = P exp(− γ(0)
Aiµ (γ(t))dxµ )

and with gi (0) = e [2].

(e) Holonomy. Then, defining a loop as, Cp M = {γ : [0, 1] → M |γ(0) =


γ(1) = p}, where u ∈ E, and π(u) = p we can define the holonomy group as,

Φu = {g ∈ G|τγ (u) = ug, γ ∈ Cp (M )}

where G is the structure group, or the Lie group, of the G-bundle.

Meanwhile, parallel transport around a loop is defined,

gγ = exp(− Aiµ dxµ )


H

Thus, as a vector in the horizontal lift is parallel transported around a very


small loop, it still of course will come back slightly rotated by an element of its
holonomy group. This rotation however, is explained in terms of the curvature,
and as a projection onto the vertical subspace, thus onto the Lie algebra [2].

(f ) Curvature. In order to explain this projection, we first we define the


curvature on a principal G-bundle as the covariant derivative of the connection
one-form, ω,

Ω = Dω ∈ Ω2 (E) ⊗ g.

Then by the Cartan structure equation, we know that Ω and ω satisfy,

Ω(X, Y ) = dp ω(X, Y ) + [ω(X), ω(Y )].

Of course, since X, Y ∈ Hu E, then by definition of vectors in Hu E, ω(X) =


ω(Y ) = 0. Thus we know that, [ω(X), ω(Y )] = 0,

thus, we have, Ω(X, Y ) = dp ω(X, Y ).

Then we have,

dp ω(X, Y ) = Xω(Y ) − Y ω(X) − ω([X, Y ])

but again since X, Y ∈ Hu E, we have,

Ω(X, Y ) = −ω([X, Y ]).

25
Thus if the curvature is nonzero, that is, if Ω(X, Y ) is nonzero, then −ω([X, Y ])
is also nonzero. But if ω([X, Y ]) is nonzero, then [X, Y ] is not an element of
the horizontal subspace, Hu E. Thus, the curvature, or the failure of vectors to
close upon parallel transport around a loop, is measured by the a projection
onto the vertical subspace, Vu E, which is an element of the Lie algebra, g.

When parallel transporting vectors in the horizontal lift, X, Y ∈ Hu E around


an infinitesimal parallelogram of side length , the failure of the horizontal lift
vector to return to its original position, is a projection onto the vertical subspace,
Vu E, that is, onto the Lie algebra.

This is shown as follows,

∂ ∂
π∗ ([X, Y ]H ) = δ[V, W ] = δ[ ∂x 1 , ∂x2 ] = 0,

where we let π∗ X = V and π∗ Y = δW .

Thus, the horizontal component of Lie bracket of the vectors is zero, that is,
the Lie bracket of the two vectors, [X, Y ] is vertical.

The curvature measures this projection onto the vertical, and is an element
of the Lie algebra.

Ω(X, Y ) = −ω([X, Y ]) = A,

where [X, Y ] ∈ A] , and A ∈ g, [2].

Hence, the Ambrose-Singer theorem,

(g) The Ambrose-Singer theorem. Let P (M, G) be a G bundle over a con-


nected manifold M . The Lie algebra h of the holonomy group Φu0 of a point
u0 ∈ P agrees with the subalgebra of g spanned by the elements of the form,

Ωu (X, Y ) for X, Y ∈ Hu E

where u ∈ E [2].

Thus again, our curvature of spacetime, on an infinitely small loop, is an


element of the Lie algebra so(3, 1). Of course, we also still have our isomorphism
with the Lie algebra sl(2, C).

This curvature does however, of course, apply more broadly to other forces
and models in physics, in particular to Yang-Mills theory.

26
C. Conserved quantities, symmetry, and gauge theory.

Since in our new framework we are asserting that spacetime is perhaps curved
around all objects, this curvature may thus also correspond to a conserved
quantity, for any and all objects, regardless of size.

As well, this potentially conserved quantity points to a symmetry in the sense


of Emmy Noether’s theorem [8]. In our framework, the curvature of spacetime
around an object does not change with the size of the object, however, if one
allows that spacetime acts as a sort of phenomenological lense [5], then the
object’s appearance to its observer, would change depending on the size of the
object with respect to this observer.

Loosely speaking and needing further specification, one could apply these
ideas of symmetry and conserved quantities to consider the Lagrangian for the
Yang-Mills equation [1],

L = 12 tr(F ∧ ?F ) = 14 tr(Fµν F µν )vol.

First, since we have this isomorphism between Hermitian 2 x 2 matrices and


Minkowski spacetime, such that an element of SL(2, C) can be written, X(x) =
xµ σµ and an element of Minkowski spacetime can be written, xµ = 21 tr(σµ X),
then our Yang-Mills Lagrangian could be written in either framework, where
Fµν represents the curvature of spacetime.

Hermann Weyl’s work on gauge theory, which was the future basis for Yang-
Mills theory, was inspired by the work on symmetry and conserved quantities
from Emmy Noether [8] [13]. It does seem that the key to uniting the two ge-
ometries of the general theory of relativity’s pseudo-Riemannian geometry, and
the Ehreshmannian geometry of the Yang-Mills theory, is through reinterpret-
ing the general theory of relativity and its relationship to quantum mechanics,
within the symmetries of this new framework.

27
IV. CONCLUSION

It has been shown in this paper shown how one can flip physics upside down,
from a world of small, medium, and large objects observed by ourselves, to a
world instead of small, medium, and large observers of any and all objects.

I have based this proposal on the fact that size is not an intrinsically mean-
ingful property. The size of an object only becomes meaningful relative to the
size of its observer. I have also based this proposal on an extension of Einstein’s
equivalence postulate. Einstein’s equivalence postulate states that it is impos-
sible for an observer to tell the difference, in theory, between an accelerating
reference frame, and being acted upon by a gravitational force. Here, it can be
said that a large observer of Jupiter’s size looking at us, is theoretically, and
with more exact number size ratios, indistinguishable from one of us looking at
a proton. The converse is also true. For either of these observers, it would be
impossible to discern a difference between these two situations.

I have also, in this proposal, taken a contrarian point of view, namely, that
it is in fact not quantum mechanics, but the general theory of relativity that is
actually the more fundamental theory of the two. The first reason for this is that
the realm of the small objects taken to define the realm of quantum mechanics,
only applies as a meaningful concept to us. An object being small, relative to
the size of a human being, does not make it inherently more fundamental to
physics, rather only from our perspective. However, the fabric of spacetime is
an intrinsically meaningful concept that contains and transcends any objects
within it. It is most fundamental of all to an observer of physics, from any
perspective and it is unavoidable in the human experience of the universe at all
levels. The second reason for taking general relativity as our starting point is
that it leads to less ridiculous conclusions and possibly sheds light and clarity on
quantum mechanics’s mysterious nature. The curvature of spacetime, from our
much larger perspective, may in fact be what causes the appearance of quantum
mechanics.

As a consequence, this proposal removes much of the major philosophical and


experimental dichotomy between the general theory of relativity and quantum
mechanics. The general theory of relativity is no longer just the theory of large
objects, but can instead be thought of as the theory of all objects, when the
observer is sufficiently small by comparison. Likewise, quantum mechanics need
no longer only be the theory of very small objects, but instead can be thought of
as the theory of any object, when perceived by a sufficiently large observer. They
are no longer two major theories that apply to disparate realms of the universe,
but rather can be thought of as applying at all times, and simultaneously, to all
objects in the universe.

Mathematically, we started with a large object like Jupiter and allowed for

28
it to be following the general theory of relativity and thus curving spacetime
around itself. Then, instead of shrinking the object, we made much larger
its observer, so that Jupiter had the apparent size of a proton. As Jupiter
progressed along a path in spacetime, we showed that this equation of motion
is at least locally isomorphic to the appearance of the proton’s motion in the
mathematical setting of quantum mechanics. Thus Jupiter, from the perspective
of a much larger observer, seems to mimic the mathematical appearance of the
motion of a proton in quantum mechanics.

By further philosophical and mathematical extension, we have shown first


how in addition to the Lie algebra isomorphism between so(3, 1) and sl(2, C),
that there is also the structure appropriate for a Lorentz transformation us-
ing an action of SL(2, C). Second, we have shown how through our inverted
framework, one could more suitably allow for an alternative principal G-bundle
structure for the general theory of relativity. This new structure could more
closely resemble and accommodate Yang-Mills theory, perhaps, as the inverse
of quantum mechanics, and thus also as an underlying framework for the other
fundamental forces. Third, thinking loosely along the lines of Emmy Noether’s
theorem, perhaps using the curvature of spacetime around objects as a new con-
served quantity and invariance in size differences as a new symmetry, one could
perhaps construct our underlying framework of physics from this new symme-
try. Within this framework, the laws of physics do not change when the size
differences shift over objects of different sizes, rather the laws of physics remain
invariant. It is through this inverted framework, and in coalescence with the
ideas of Emmy Noether, Hermann Weyl, Albert Einstein, Chen Ning Yang and
Robert Mills, that we are propelled forward into a new underlying structure.

Thus, we have shown how, philosophically, experimentally, and mathemati-


cally, one could resolve the dichotomy between the general theory of relativity
and quantum mechanics. We have a theory of physics that has been flipped up-
side down from small, medium, and large objects, to small, medium, and large
observers, an Inverse Theory of Relativity. Perhaps, in accord with this theory,
Herr Professor Doktor Einstein was correct after all, when he said, ”God does
not play dice” [11].

29
In Memoriam: James Orlan Skelton

With heartfelt thanks to my Mom, Helen Talei Skelton, for, everything.

In Pectore:

With heartfelt thanks to Dr. Jeff Wragg from the University of Charleston, for
being willing to read my paper and offer excellent criticism, questions, and
advisement, and also for being such a wonderful teacher many years ago.

With heartfelt thanks to Andrew Sturges for his love, wisdom, and
encouragement.

30
References
[1] John Baez and Javier P. Muniain, Gauge Fields, Knots and Gravity (World
Scientific, Singapore, 1994), Vol. 4, pp. 231 - 238, pp. 243 - 247, p. 371 - 381.
[2] Mikio Nakahara, Geometry, Topology, and Physics (IOP Publishing, Bristol,
1990), pp. 178 - 179, pp. 232 - 233, pp. 329 - 343.
[3] T.L. Curtright, Z. Cao, A. Peca, D. Sarker, and B.D. Shrestha, Lie Groups
and Propagators Exemplified arXiv: 2112.14401.
[4] Steven Weinberg, Foundations, The Quantum Theory of Fields (Cambridge
University Press, Cambridge, 1995), Vol. 1, p. 61.
[5] Edmund Husserl, The Paris Lectures, translated by Peter Koestenbaum,
(Kluwer Academic Publishers, Boston, 1998), pp. XXXI - XXXIV, p. XLIII
- XLIV.
[6] Florian Conrady and Jeff Hynbida, Unitary irreducible representations of
SL(2,C) in discrete and continuous SU(1,1) bases arXiv:1007.0937v2.
[7] Immanuel Kant, Critique of Pure Reason, translated and edited by Paul
Guyer and Allen W. Wood (Cambridge University Press, Cambridge, 1998),
p. 10.
[8] Dwight E. Neuenschwander, Emmy Noether’s Wonderful Theorem (John
Hopkins University Press, Baltimore, 2017), pp. 8 - 9, 84 - 85, 186, 234.

[9] Robert M. Wald, General Relativity (University of Chicago Press, Chicago,


1984). pp. 8 - 9, 36 - 38, 66-67, 73-74, 77.
[10] The Nobel Prize in Physics, Nobelprize.org. Nobel Media AB (2014). Web.
(3 Jul 2018).
[11] Andrew Robinson, Did Einstein really say that? Nature. 557, 30 (2018).

[12] R. Shankar, Principles of Quantum Mechanics (Kluwer Academic / Plenum


Publishers, New York, 1994), 2nd edition, p. 122 - 123, 237.
[13] Jim Baggott, The Quantum Story, A History in 40 Moments (Oxford Uni-
versity Press, 2011), pp. 193 - 202.

[14] Roger Scruton, Kant, A Very Short Introducction (Oxford University Press,
2001), pp. 47 - 48.

31

You might also like