You are on page 1of 57

Yang-Mills theories and confinement

vs.
Gravity and geometry∗
Antônio D. Pereira,1, † Rodrigo F. Sobreiro,1, ‡ and Anderson A. Tomaz1, §
1
UFF − Universidade Federal Fluminense,
Instituto de Fı́sica, Campus da Praia Vermelha,
Avenida General Milton Tavares de Souza s/n, 24210-346,
Niterói, RJ, Brasil.
(Dated: July 2015)
These lectures contains a dense material about gauge theories and the first order formalism of
gravity. A scenario where quantum gravity is realized as a Yang-Mills theory is discussed with some
detail. In this scenario, gravity as a geometrodynamical effect is an emergent phenomenon.

FT
RA
D

∗ Lectures presented at the X Escola do Centro Brasileiro de Pesquisas Fı́sicas. 13-24 July 2015.
† Electronic address: aduarte@if.uff.br
‡ Electronic address: sobreiro@if.uff.br
§ Electronic address: tomaz@if.uff.br
2

Contents

INTRODUCTION 4

PART I: GAUGE THEORIES AND THE GRIBOV PROBLEM 5

I. Gauge theories and the gauge principle 5


A. Electromagnetism 5
B. Covariant formulation and U (1) symmetry 7
1. Definitions 7
2. Maxwell equations and gauge fixing 8
C. Non-Abelian gauge theories 9
1. Lie groups basics 9
2. Gauge field, field strength and Yang-Mills action 11

II. Quantization of gauge theories 12


A. Faddeev-Popov quantization 13

FT
B. BRST quantization 15
C. Renormalization, physical states and unitarity 16
D. Asymptotic freedom and confinement 17

III. Gribov ambiguities 18


A. Statement of the problem 18
B. Gribov’s solution 19
C. The Gribov-Zwanziger action and BRST soft breaking 22
D. The Refined Gribov-Zwanziger action 23

PART II: GRAVITY AS A GAUGE THEORY 25


RA
IV. Equivalence principle and geometrodynamics 25
A. General relativity in the metric formalism 25
B. First order formalism of gravity 27
1. Vierbein 27
2. Spin-connection 28
3. Curvature, torsion and hierarchy identities 29
4. Einstein-Hilbert action 30
C. Gravity as a gauge theory 31

V. Generalized theories 31
A. Lovelock theory 31
D

B. Mardones-Zanelli theory 32

VI. Quantization attempts 33

PART III: EFFECTIVE GRAVITY FROM YANG-MILLS THEORIES 34

VII. General ideas 34

VIII. SO(5) Yang-Mills theory 36


A. Projection and rescaling 36
B. Contraction and symmetry breaking 37

IX. Induced gravity 38

X. Quantum aspects 38
A. Weinberg-Witten theorems, emergent gravity and spin-1 gravitons 38
B. Cosmological constant problem 39
C. Quantum predictions 39
1. Running parameters 40
2. Numerical estimates 41
3

XI. Classical aspects 42


A. Classical field equations 42
B. Spherical symmetry static solutions 43
C. Cosmology 44
1. The Λ-CDM model 45
2. High curvature regime 45
D. Dark stuff 48

A. Differential forms 48
1. Exterior product and p-forms 48
2. Exterior derivative 49
3. The Hodge dual operator 49
4. Integration of differential forms 50

B. Geometrical aspects of gauge theories 50


1. Principal bundles and connections 51
2. Universal bundles 52
3. Gravity geometrical structure 53

FT
Acknowledgments 54

References 54
RA
D
4

INTRODUCTION

Gauge theories[1–6] have been extremely successful in describing at least three of the four known fundamental
interactions in Nature, at classical and quantum level. Except for gravity, the Standard Model is the state of the art
of quantum field theory and particle physics where Yang-Mills theories [7, 8] is the main framework. Gravity, on the
other hand, is not described by interacting fields but by a geometrodynamical theory for the spacetime itself [9–13].
In despite of the success of both type theories in describing Nature, some open issues remain to be understood. In
particular, from the theoretical point of view, confinement and quantum gravity are, perhaps, the most important
open problems of these theories.
Confinement refers to a specific behavior of the strong sector of the Standard model, known as Quantum Chromo-
dynamics (QCD) which describes, essentially, the dynamics of quarks and gluons. The SU (3) Yang Mills (YM) gauge
theory perfectly describes this dynamics at high energies where quarks and gluons are almost free. It displays the so
called asymptotic freedom characterizing the increasing of the coupling parameter as the energy decreases. Thus, as
lower the energy is, as higher the force between quarks and gluons is. The low energy sector is highly non-perturbative
and very difficult to be handled. Nevertheless, V. N. Gribov has found in in the late seventies [14, 15] that, at this
region, the theory lacks of a consistent quantization at the non-perturbative level. This is a technical problem, known
as Gribov ambiguities, which must be considered for quantum consistency. Remarkably, the improvement of the

FT
quantization leads naturally to confinement evidences in a highly non-trivial way.
On the other hand, gravity, as a geometrodynamical phenomenon, lacks for a quantum version (even perturbatively).
In many aspects, the principles of general relativity seem to be incompatible to the principles of quantum field theory,
dooming their union to failure. Many alternative approaches to describe gravity at quantum levels have been proposed
for almost one century. Widely known examples are string theory [16–19], loop quantum gravity [20, 21], asymptotic
safety [22, 23], Hořava-Lifshitz gravity [24–26], the vast scenarios of emergent gravities [27–40] and so on. In particular,
emergent gravities possibly solve the problem by stating that gravity as a geometrodynamical phenomenon can not
be quantized. Gravity should be an emergent feature of a more fundamental theory that have nothing to do with the
spacetime dynamics at quantum level.
In these lectures we discuss an interesting analogy between non-perturbative QCD and gravity [37–42]. In this
RA
proposal, gravity emerges from a Yang-Mills theory for the gauge symmetry group SO(5) in four dimensions. Thus, we
will discuss the details of Yang-Mills theories and the Gribov problem as well as the first order formalism of gravity in
order to reach the goal of understanding how gravity can emerge from a quantum gauge theory. Specifically, quantum
gravity will be described by a Yang-Mills action in flat space while classical gravity will be obtained from it as a
geometrodynamical theory for the spacetime and both sectors should not intersect. Before that, a dense discussion
about Yang-Mills theories and the first order theory of gravity are discussed.
The idea of these lectures is to provide a complete view of gauge theories and gravity in a conceptual way. Due
to the vast amount of results known in these theories, many details are omitted in favor of conceptual discussions.
A lot of technical passages are left as exercises of all kinds of difficulty levels. These exercise are recommended for
students interested in explore the field. Previous notions in quantum field theory (QFT), differential geometry, group
theory and general relativity is advisable, but not mandatory. Along the text many references are suggested and two
appendices (about differential forms and fibre bundles) is included for self-consistency. At the end of the course, the
D

student should acquire a lot of conceptual information about current research on gauge theories, confinement, gravity
and quantum gravity.
5

PART I: GAUGE THEORIES AND THE GRIBOV PROBLEM

In this part we focus on the construction of gauge theories. Starting with Maxwell equations as an Abelian
gauge theory, we define gauge field, gauge transformations and the gauge principle. A few concepts in Lie groups
and Lie algebras are introduced in order to discuss non-Abelian gauge theories, which are the main theme of this
part. Quantization is discussed with some detail within Faddeev-Popov’s method. Then, BRST symmetry and
its consequences, including BRST quantization, are carried out. The concepts of physical states, renormalizability,
unitarity, asymptotic freedom and confinement are established. All these aspects are mainly found in quantum field
theory text books, see for instance [1–4, 6, 43]. The issue of Gribov ambiguities is then addressed: a review since
Gribov‘s seminal work back in 1978 is compiled with the most important results and concepts. The main references
about the Gribov problem are [14, 15, 44–50].

I. GAUGE THEORIES AND THE GAUGE PRINCIPLE

A. Electromagnetism

FT
The most known example of a gauge theory is, perhaps, electromagnetism [1, 51, 52]. In its non-covariant form,
electromagnetism is described by Maxwell equations in vacuum,
ρ
∇·E = ,
o
∇·B = 0,
∂B
∇×E+ = 0,
∂t
∂E
∇ × B − o µo = µo j , (1)
∂t
RA
where E and B are the electric and magnetic fields, respectively; o is the electric permittivity of the vacuum and µo
is the magnetic permeability of the vacuum; the charge density is ρ while j = ρv is the current density with velocity
v.
Maxwell equations can be simplified by the introduction of potential fields. In fact, the second equation (magnetic
Gauss law - first homogeneous equation) states that a solution for B is

B=∇×A, (2)

where A is the vector potential field. Using this solution at Faraday law’s (second homogeneous equation) it is possible
to show that the solution for the electric field is1
D

∂A
E = −∇φ − , (3)
∂t
where φ is the electric scalar potential. Remarkably, the vector and scalar potentials can always be redefined in such
a way that the electric and magnetic fields are left invariant (and so are the Maxwell equations, obviously). These
redefinitions are called gauge transformations:
∂ξ
φ 7−→ φ − ,
∂t
A −
7 → A + ∇ξ , (4)

with ξ = ξ(r, t) being the gauge transformation parameter. As opposed to coordinate transformations (for example,
Lorentz transformations), gauge transformations are transformations that leave spacetime coordinates intact, it is a
type of transformation that affects only fields. The fields φ and A are recognized as the gauge fields. The electric and

1 The fact that the homogeneous equations are used to define the potentials are not accidental. These equations have a deep geometrical
meaning, which we will briefly discuss in this section and along this part.
6

magnetic fields are said to be gauge invariants. In terms of the gauge fields, the Maxwell equations (1) read (Exercise
1.1)
∂ ρ
−∇2 φ −
∇·A = ,
∂t o
∂2A
 
∂φ
−∇2 A + o µo 2 + ∇ ∇ · A + o µo = µo j . (5)
∂t ∂t
We notice that the homogeneous equations are automatically satisfied through (2) and (3). Thus, (5) are obtained
from the substitution of these solutions into the inhomogeneous Maxwell equations. Although equations (5) appear
complicated with respect to the original equations (1), it is important to stress out that E and B carry six components
of freedom while φ and A carry only four.
The fact that the gauge fields can be redefined up to a scalar function establishes that they are not observables
because there is an infinite number of equivalent fields leading to the same pair of electric and magnetic fields,
recognized as the actual physical fields. Thus, no matter how one chooses ξ to solve a problem, the corresponding
electric and magnetic fields will always be the same. Moreover, to solve a specific problem, it is possible to choose a
unique class of gauge fields in order to simplify the equations. This procedure is called gauge fixing and is performed
by the introduction of a constraint for the gauge potentials. For instance, the Coulomb gauge is characterized by

FT
transverse vector potentials, ∇ · A = 0, leading to
ρ
−∇2 φ = ,
o
∂2A ∂
−∇2 A + o µo 2 = µo j − o µo ∇φ , (6)
∂t ∂t
Then, the scalar potential is the solution of the Poisson equation and is an instantaneous field, a non-physical feature,
while the vector potential is the solution of a inhomogeneous wave equation. Another example (Exercise 1.2) is the
Landau gauge2 , ∇ · A + o µo ∂φ
∂t = 0, which leads to inhomogeneous wave equations for both, φ and A.
The difference between physical and non-physical quantities can be summarized by the so called gauge principle
RA
Gauge Principle: Physical observables are described by gauge invariant quantities.

Since E and B are gauge invariants, anything constructed with them are also gauge invariant, for instance,
the electromagnetic energy density o E 2 /2 + B 2 /2µo . In this way, although physical, neither E or B are fundamental.
The fundamental fields are φ and A, in the sense that they are “indivisible”. Hence, a gauge theory is constructed
from the gauge fields (as fundamental objects) of the theory. However, due to the gauge symmetry, they can not be
observables. Instead, any observable must be constructed from gauge fields and has to be gauge invariant.
A very beautiful example (Exercise 1.3) of the gauge principle is the Aharonov-Bohm effect [53, 54]. Let us consider
a long thin solenoid with constant magnetic field given by B = Bo k̂ inside the solenoid and 0 otherwise. A vector
potential whose curl produces such magnetic field is A = Φ/(2πr)ϕ̂ where ϕ̂ is the polar angle in cylindrical coordinates
D

and
Z
Φ= B · dS . (7)
S

is the magnetic flux through a surface whose boundary is a curve around the solenoid. Then, an electron beam is
split in two, with each sub-beam passing on different sides of the solenoid (This can be done by the the double-slit
apparatus). It can be shown that the interference pattern between both sub-beams on a screen will be shifted by a
factor
 I 
e
δ = exp A · dl . (8)
~ C
with respect to the case with no solenoid. In (8), e is the electron charge and ~ is the Planck constant. Since this is
an observable phenomenon, it should be consistent with the gauge principle. Indeed, (8) is gauge invariant:
I I I
e e e
A · dl 7−→ A · dl + ∇ξ · dl . (9)
~ C ~ C ~ C

2 Also known as Lorentz gauge.


7

Since the last term vanishes identically, the shift (8) is gauge invariant. Moreover, due to Stokes theorem
Z Z
ln δ = ∇ × A · dS = B · dS = Φ , (10)
S S

where S is any surface with boundary C. Thus, the shift of the interference pattern, although given in terms of A, is
a gauge invariant quantity. This illustrates the fact that, even though A is not physical, it can generate observable
phenomena through gauge invariant objects constructed from it and starting from postulates which are based on
gauge concepts.

B. Covariant formulation and U (1) symmetry

We will now be a bit more formal in constructing the classical electromagnetism as a gauge theory by writing it
down in a manifestly covariant form, since it is a natural relativistic theory.

1. Definitions

FT
We define Greek indexes as collective indexes for space and time, α, β, . . . ∈ {0, 1, 2, 3}, where the 0th coordinate
stands for time. The derivative operator will be written in a simplified form as ∂µ = ∂/∂xµ . Moreover, from now on,
we adopt natural units3 and the 4-dimensional Euclidean4 metric will be assumed, i.e., δµν ≡ diag(1, 1, 1, 1).
We consider a four dimensional gauge field Aµ and its gauge transformation
1
Aµ 7−→ Aµ + ∂µ ξ . (11)
e
This definition coincides with the gauge transformations (4) for A0 = φ (Exercise 1.4) and the factor 1/e is introduced
for convenience. The gauge transformation (11) is a gauge transformation of a gauge connection under the action of
RA
an element of the U (1) group, i.e., the group of complex numbers with unit modulus. In fact, (11) can be rewritten
as
1
Aµ 7−→ Aµ + e−ξ ∂µ eξ . (12)
e
where5 eξ ∈ U (1).
Formally, the gauge field is a connection over an abstract topological space6 which we will call gauge space. Thus,
a covariant derivative can be defined as
Dµ = ∂µ − ieAµ . (13)
where e is the coupling parameter which is typically associated with the electron charge in QED. In general, covariant
derivatives do not commute and define the curvature of the gauge space (commonly known as the field strength)
D

(Exercise 1.5):
[Dµ , Dν ] = −ieFµν = −ie (∂µ Aν − ∂ν Aµ ) . (14)
where the derivatives are supposed to act over a generic charged field, for instance, a spinor field describing the
electron. It is simple to show that the field strength is gauge invariant (Exercise 1.6). The tensor Fµν is obviously
anti-symmetric. Thus, it can be described by an anti-symmetric 4 × 4 matrix whose components coincide with the
electric and magnetic fields (Exercise 1.7),
0 E1 E2 E3
 
−E1 0 B3 −B2 
Fµν ≡  (15)
−E2 −B3 0 B1 
−E3 B2 −B1 0

3 From now on, unless the contrary is said, we will employ natural units c = ~ = 1.
4 Minkowskian metric would be preferable for perturbative treatments. However, we will deal with non-perturbative phenomena whether
Wick’s rotation is not established. Thus, it is often chosen to start from Euclidean metric, where we can actually perform explicit
computations in quantum field theory.
5 Any complex number with unit norm can be written in this form.
6 See Ap. B.
8

The field strength is clearly gauge invariant, a property which makes it an observable of the theory, as it should be
due to its gauge invariant components.

2. Maxwell equations and gauge fixing

In the covariant formulation, Maxwell equations (1) read

∂ν Fµν = jµ ,
µναβ ∂α Fµν = 0 , (16)

where jµ is the four-current density, jµ ≡ (ρ, ρv). The first set of the Maxwell equations above coincide with the
inhomogeneous equations of (1) (Exercise 1.8) while the second set coincide with the homogeneous equations of (1)
(Exercise 1.9). In fact, analogously to (2) and (3), the solution of the second equation of (16) is (14). Moreover,
this is a topological equation because the solution is the definition of the gauge potential, i.e. the definition of the
gauge potential automatically satisfies this equation. These equations are recognized as the electromagnetic Bianchi
identities.

FT
Maxwell equations (16) can be rewritten in terms of the gauge field (Exercise 1.10): the first of (16) leads to

∂µ ∂ν Aν − Aµ = jµ , (17)

while the second is automatically satisfied, as discussed.


A useful way to simplify the equations (17) is to fix a gauge, which means that we can impose a constraint over Aµ
in order to select a class of gauge potentials among all possible gauge potentials that could lead to a consistent field
strength solution. Typicality, the linear covariant gauges are imposed: ∂µ Aµ = f , where f is a spacetime function.
The special case f = 0 is recognized as the Landau gauge while the non-covariant gauge ∂i Ai = 0 is the Coulomb
gauge. In the case of the Landau gauge, for instance, Maxwell equations (17) simplify to
RA
Aµ = −jµ , (18)

It is very useful to define an action for electromagnetism. A typical choice is


Z   Z
4 1 α 
SM = d x Fµν Fµν − jµ Aµ + d4 x b2 + b∂µ Aµ , (19)
4 2

where b is the Lautrup-Nakanishi field [43] which is a Lagrange multiplier enforcing the linear covariant gauge fixing.
It is a straightforward exercise (Exercise 1.11) to show that the minimization of this action with respect to Aµ leads
to a gauge fixed Maxwell equations and the minimization with respect to b leads to the gauge fixing condition7 .
In summary, electrodynamics, as described by the action (19), can be obtained by postulating:
D

• Gauge principle with respect to the U (1) group - This postulate naturally leads to the existence of a gauge
connection and a gauge curvature.
• (Euclidean) Lorentz invariance of the action - This postulate ensures the principle of relativity. Moreover,
Lorentz transformations form a group whose Euclidean version is the group of four-dimensional rotations in
spacetime, the SO(4) group.
• Further requirements as locality8 and absence of higher order derivatives can be employed in a consistent way
in order to avoid unwanted terms in (19).
• Any field coupled to Aµ must respect all above postulates.

7 In field theory, the fields equations can be obtained through the generalized Euler-Lagrange equations:
 
∂L ∂L
− ∂ µ =0.
∂ΦA ∂ (∂µ ΦA )

where ΦA is a generic field with A being a collective index.


8 Roughly speaking, a local quantity in QFT is an object that only depends on a single spacetime point.
9

C. Non-Abelian gauge theories

From now on, due to the intricacies of non-Abelian theories (which will become evident as we proceed), we will
neglect the coupling with external sources jµ or other fields.

1. Lie groups basics

Our aim in this section is to provide a brief overview of what we need about Lie groups and group theory. It is not
our intent to provide detailed discussion on this matter. For detailed texts about Lie Groups, see for instance [55–57]
and also [1] for an extended, yet summarized, introduction. To start with, let us enunciate the axioms of a group.
Consider a set G of elements u and an operation ·, then G is said to be a group if its elements obey:
• Closure: If u1 and u2 ∈ G then u3 = u1 · u2 ∈ G.

• Associativity: If u1 , u2 and u3 ∈ G, then (u1 · u2 ) · u3 = u1 · (u2 · u3 ).

FT

• Identity: ∃ uo = 1 ∈ G u · 1 = 1 · u = u ∀u ∈ G.


• Inverse: ∃ u−1 ∈ G u−1 · u = u · u−1 = 1 ∀u ∈ G.

Moreover,
• If G is infinite and continuous, the group is said to be a Lie Group.
• If [u1 , u2 ] = 0 ∀ u1 and u2 ∈ G, the group is said to be an Abelian group, otherwise it is said to be a non-Abelian
RA
group.
The notion of subgroup will also be important along this lectures:
• If G is a group and H ∈ G is a smaller set than G, then H is a subgroup of G if H is also a group, i.e. if the
elements of H obeys the four group postulates.
Lie groups are typically, but not exclusively, N × N matrix groups (the case N=1 is a group formed by scalars, like
U (1)). A useful property of Lie groups is that their elements can always be written as exponentials u = eξ , where ξ
is the gauge parameter. When the gauge parameter is a function of the spacetime coordinates, the group is said to
be a local group, otherwise the group is said to be a global group. The algebra of the group is the tangent space of
the group at the identity[1]. For any element near the identity, we can write u ≈ 1 + ξ. We can expand ξ through
a matrix basis T A as ξ = ξ A T A where the capital Latin indexes A, B, . . . , H ∈ {1, 2, . . . , dim G}. The exponent ξ
D

is said to belong to the algebra of G. It should be clear that if · is multiplication, its action with respect to ξ and
T A is addition. It is important to have in mind that, in general, the dimension of the group does not coincide with
the dimension of its elements, dim G 6= N . The basis elements T A are called generators. The characterization of the
algebra is given by the commutation rules of the generators,
[T A , T B ] = f ABC T C , (20)
where the skew-symmetric object f ABC is composed by the structure constants of the group. If f ABC = 0, then the
group is obviously Abelian. Moreover, the structure constants obey the Jacobi relations (which can be derived from
(20))
f ABC f CDE + f ADC f CEB + f AEC f CBD = 0 . (21)
As a supplementary condition, the generators are normalized as
1
Tr(T A T B ) = − δ AB . (22)
2
A representation of a group is the realization of the elements u ∈ G as a linear operator U (u) acting on a linear
space. Thus, given the elements of V , denoted by ψ i , a representation is characterized by
i
ψ 0 = U ij ψ j . (23)
10

It is demanded that

U (u1 u2 ) = U (u1 )U (u2 ) ,


U (u−1 ) = [U (u)]−1 ,
U (1) = 1 . (24)

Clearly, this is a mapping u 7−→ U which will be reflected on the generators of G. Hence, (24) implies that

U (T A + T B ) = U (T A ) + U (T B ) ,
U (aT A ) = aU (T A ) ,
U ([T A , T B ]) = [U (T A ), U (T A )] , (25)

where a ∈ R. Thus, we also have

U (u ≈ 1 + ξ) ≈ 1 + U (ξ) = 1 + ξ A U (T A ) . (26)

In gauge theories, the most relevant representations are the fundamental representation and the adjoint represen-

FT
tation. For the fundamental representation, let us consider G as a group of N × N matrices. Then, an N -dimensional
I
vector space V with elements ψ I , transform under G as ψ 0 = U IJ ψ J , where I, J, K, . . . ∈ 1, 2, 3, . . . N . In this case,
the elements U are simply N × N matrices.
The adjoint representation is a direct realization of the algebra of G. The action of an element u in the adjoint
representation on an element of the algebra ξ is given by Ad(u)ξ = u−1 ξu. Then, u−1 ξu is required to belong to the
adjoint representation as well. For u ≈ 1 + η it is then clear that (Exercise 1.12) Ad(η)ξ = [ξ, η]. Thus, if ξ and η are
simply the generators, we have that Ad(T B )T A = f ABC T C . On the other hand, the definition of a representation
(23) yields

[T A ]BC = f ABC , (27)


RA
which means that in the adjoint representation the dimension of the elements coincide with the group dimension. It
is a straightforward exercise (Exercise 1.13) to check that both, fundamental and adjoint, representations respect (24)
and (25).
Let us discuss a few important groups. We already discussed the unitary Abelian group U (1) which is the group of
all complex numbers with unitary modulus, i.e. zz ∗ = 1. Taking z = eiξ , where ξ ∈ R, it is almost trivial to check all
four group axioms (Exercise 1.14).
The generalization of the unitary group to matrix groups is realized by the group of unitary N ×N complex matrices,
i.e. uu† = 1. This group is denoted by U (N ). From its definition we have that det u = ±1 and dim U (N ) = N 2 − 1.
The restriction for positive determinants excludes discrete transformations such as reflections. This is the so called
special unitary group, denoted by SU (N ). It is easy to verify that unitary groups respect the group axioms (Exercise
1.15). Geometrically, the elements u ∈ SU (N ) are rotations of complex vectors.
D

The group of N × N orthogonal real matrices is denoted by O(N ) and obeys uuT = 1 where uT is the transpose
of u. Obviously, det u = ±1. It is clear that u ∈ O(N ) represent rotations and reflections of real vectors in an
N -dimensional Euclidean space and hence dim O(N ) = N (N − 1)/2. The restriction to pure rotations leads to
the special orthogonal groups SO(N ). Again, the verification of the group axioms is very easy (Exercise 1.16). The
algebra of SO(N ) is related to the Euclidean metric. In fact, in the adjoint representation, the generators are matrices
anti-symmetric in their indexes, T ab = −T ba , where a, b, c, . . . , h ∈ {1, 2, 3, . . . , N (N − 1)/2}, obeying

[T ab , T bc ] = f abcdef T ef ,
1  ad be
f abcdef = − δ δ − δ ae δ bd δ cf + δ ac δ be − δ ae δ bc δ df .
  
(28)
2
All examples mentioned above assume · to be matrix multiplication. The group of translations, commonly denoted
by RN , is an example of a group whose operation is addition. Hence, given an N -dimensional vector space V , the
elements u ∈ RN are also N -dimensional vectors in the fundamental representation. The action of the group elements
on the elements of V is given by ψ 0 = ψ + u. In this case, the identity is the null vector ψoi = 0 and the inverse is
−u. Closure and associativity can be directly verified (Exercise 1.17). Because translations commute, the algebra of
translations is characterized by

[P A , P B ] = 0 . (29)
11

It is possible to define a more general group by considering rotations and translations in a single larger group, the
so called Euclidean Poincaré groups ISO(N ) = SO(N ) n RN , where n is a semi-direct product9 . The action of the
Euclidean Poincaré group on a vector ψ ∈ V is

ψ 0 = Λψ + u , (30)

where ψ is written as a N -dimensional column vector, Λ = Λab T ab ∈ SO(N ) and u = ua P a ∈ RN . To show that
ISO(N ) is a group is also an easy exercise, yet bigger than the previous ones (Exercise 1.18).
In the same way, the properties of the Lorentz group SO(1, 3) (which is the group of four-dimensional orthogonal
matrices preserving the Minkowskian metric) and the general linear group GL(N, R) (which is the group of all N × N
real matrices) can be easily checked (Exercise 1.19).
In what follows, we will stick mainly to Lie groups which are semi-simple and compact10 .

2. Gauge field, field strength and Yang-Mills action

The generalization from QED to non-Abelian gauge theories relies on its extension to a non-Abelian group, G.
Following the Abelian case, we start by defining the gauge field as an algebra valued connection of the group space

FT
A †
in the adjoint representation, Aµ = AA A
µ T . The generators are chosen to be anti-hermitian, (T ) = −T , so
A
A B ABC C
[T , T ] = f T . The respective covariant derivative which acts on objects with adjoint group indexes is

DµAB = δ AB ∂µ − κf ABC AC
µ , (31)

where κ is the coupling parameter (analogous to e in QED). Equation (31), in a component independent notation,
can be written as

Dµ = ∂µ + κ[Aµ , ] . (32)
RA
It is easy to show that (32) reduces to (31) (Exercise 1.20). The gauge transformations are now given by
 
1
Aµ 7−→ u−1 Aµ + ∂µ u , (33)
κ

where u ∈ G. It is convenient to consider infinitesimal transformations through u ≈ 1 + ξ, which gives (Exercise 1.21)

1 AB B
AA A
µ 7−→ Aµ + D ξ . (34)
κ µ

From (33) and (34) we can see that gauge transformations are no longer linear as in the Abelian case. This non-linear
character is the main reason why non-Abelian theories are much more complicated than Abelian ones. Again, by
D

commuting covariant derivatives, one finds the field strength (curvature of the gauge space):

[Dµ , Dν ] = κFµν . (35)

where

Fµν = ∂µ Aν − ∂ν Aµ + κ[Aµ , Aν ] . (36)

Two very important observations are in order now: First, the field strength has a non-linear piece, not present at the
Abelian case. This feature means that the gauge field interacts with itself, for instance, gluons interact among each
other (a property not observed in photons). Second, the field strength is not a gauge invariant object (Exercise 1.22),

Fµν 7−→ u−1 Fµν u . (37)

9 The semi-direct product is a direct product for which one sector is a normal subgroup, which is the case of RN . See also the next
footnote
10 Roughly speaking, a compact group is a bounded and closed group. Semi-simple group is a group that can be decomposed in simple
groups. Simple groups are groups with no normal subgroups, except for the identity and the group itself. Finally, a normal subgroup
N of G is characterized by g −1 N g ⊆ N . For the actual formal definitions we refer to [55–57]
12

In fact, this is a typical transformation rule of matter fields in the adjoint representation. So, the electromagnetic
analogues can not be associated with observables. The field strength can be written in components of the algebra,
A A
Fµν = Fµν T , as (Exercise 1.23)

A
Fµν = ∂µ AA A
ν − ∂ν Aµ + κf
ABC B C
Aµ Aν . (38)

To construct a consistent action, we evoke: The gauge principle for the group G; Euclidean Lorentz invariance;
Locality; Absence of higher order derivatives. It turns out that the simple generalization of (19) is indeed gauge
invariant (Exercise 1.24):
Z
1 A A
SY M = d4 x Fµν Fµν , (39)
4

where we have omitted the gauge fixed part (see next section) and a possible interacting term (as previously an-
nounced). This action is known as Yang-Mills action [6, 7]. The non-linear character of the field strength leads to
non-linear field equations which means that this kind of theories are physically richer as well as much complicated

FT
when compared to Abelian cases. In fact, variation of the Yang-Mills action with respect to the gauge field leads to
(Exercise 1.25)
A
Dµ Fµν =0. (40)

On the other hand, because of (38), a topological identity exists,


A
µναβ Dν Fαβ =0, (41)

which is recognized as the Bianchi identity and generalizes the homogenous electromagnetic field equations in (16).
RA
II. QUANTIZATION OF GAUGE THEORIES

Til now, we have defined only the classical quantities of gauge theories. To study the quantum aspects of Yang-
Mills theories we have to promote the fields to quantum operators or, equivalently, write down a consistent functional
integral of the exponentiated classical action. We will follow the second option here [6]. If we naively try to consider11
Z
Z = DAe−SY M , (42)

we will face inconsistencies right from the beginning. The main problem relies on the gauge symmetry: The functional
integral sums over all possible field configurations in order to provide a probability of an event or an expectation value.
D

However, the gauge symmetry establishes that there is an infinity number of equivalent fields related through gauge
transformations. Therefore, each field is actually been infinitely overcounted. Hence, the probabilities are not being
conserved. In fact, if one tries to compute the propagators of the theory (take the linearized version eq. (40) and
try to find the associated Green function), a divergent Green function is found (Exercise 1.26). Thus, even the free
theory is inconsistent (this is also true for QED).
The solution is to consider only one configuration among all equivalent fields (those connected through gauge
transformations) - for each infinite set of gauge fields related through gauge transformations, we must choose only
one representative. This procedure is called gauge fixing and is implemented by imposing a constraint over the gauge
field. The gauge fixing for the Abelian case was already discussed here, at classical level - It is easily generalized to
the non-Abelian case, also at classical level. At quantum level however, it is not a trivial task. The way to do this in
the path integral is through the Faddeev-Popov trick [58], which we discuss in the next section.

11 Since we are not going to perform any practical computation at this level, we omit the classical Schwinger sources [6].
13

A. Faddeev-Popov quantization

The question here is: How to introduce a gauge fixing in a consistent way in (42)? The idea is to integrate only
over fields obeying a certain constraint. Let us consider the linear covariant gauges ∂µ Aµ = f . The trick is to write
the unit in a non-trivial way
Z
1 = ∆(A) Duδ(f − ∂µ Auµ ) , (43)

where Auµ is the gauge orbit, which is defined through


 
−1 1
Auµ =u Aµ + ∂µ u , (44)
κ
while Aµ is a fixed, yet arbitrary, gauge configuration. It should be clear that the integral in (43) is gauge invariant
because of the closure property of G. This means that ∆(A) = ∆(Au ) . Therefore,
Z Z
−SY M (A)
Z = DAe ∆(A) Duδ(f − ∂µ Auµ ) , (45)

FT
From the gauge invariance of SY M , ∆(A) and DA, we can write
Z Z
u −SY M (Au ) u
Z = DA e ∆(A ) Duδ(f − ∂µ Auµ ) , (46)

Thus, performing the change of variables Auµ → u Auµ + κ1 ∂µ u−1 = Aµ , we achieve




Z Z
Z = DAe−SY M (A) ∆(A) Duδ(f − ∂µ Aµ ) . (47)
RA
It remains to determine ∆(A). This is easily done by changing the integration variable of (43) from u to f = ∂µ Auµ − f :
Z Z    
−1 δu δu
∆ (A) = Duδ(f) = Dfδ(f) det = det , (48)
δf δf f=0

which can be evaluated for infinitesimal gauge transformations u ≈ 1 + ξ (Exercise 1.27),


 
δf 1
∆(A) = det = det(∂µ Dµ ) . (49)
δξ f=0
κ
Thus
D

Z Z
−SY M 1
Z= DAe det(∂µ Dµ )δ(f − ∂µ Aµ ) Du . (50)
κ
which can be normalized to finally obtain
Z
Z= DAe−SY M det(∂µ Dµ )δ(f − ∂µ Aµ ) . (51)

This expression is supposedly free of gauge ambiguities. However, the determinant and the functional delta is not
in a form which we can recognize a gauge fixing, as in (19). In fact, these terms can be very complicated to handle
because they may ruin a simple algorithm to perform explicit perturbative computations such as the Feynman rules.
Fortunately, both terms can be easily exponetiated with the help of additional fields. Let us start with the functional
delta: Because gauge invariant quantities should not depend on f , we can integrate over f with a Gaussian weight12 ,
Z  Z   Z 
1 1
Df exp Tr d4 xf 2 δ(f − ∂µ Aµ ) = exp d4 xTr(∂µ Aµ )2 , (52)
α α

12 The sign of the Gaussian seems to be wrong. However, (22) still have to be employed, correcting the sign at the final expression.
14

where α is a non-negative number13 . It will be useful for the next sections to rewrite (52) as a normalized functional
integral of an auxiliary bosonic algebra-valued field b (Exercise 1.28),
 Z  Z
1 4 2
exp d xTr(∂µ Aµ ) ≡ Db exp [Tr (2b∂µ Aµ + αbb)] . (53)
α
The field b is recognized as the Lautrup-Nakanishi field [43]. Just like the Abelian case, it plays the role of a Lagrange
multiplier enforcing the gauge fixing.
The determinant is also simple lo exponentiate, it is a determinant with positive unit power, thus, it can be
substituted by a Gaussian functional integral over Grassmann variables [6]. Recalling that ∂µ Dµ is an dim G × dim G
matrix operator, we need a set of dim G2 independent variables, thus
Z  Z 
det(∂µ Dµ ) = DcDc exp − c∂µ Dµ c , (54)

where c and c are independent Grassmann variables.


The local version of (51) is then
Z
Z = DADbDcDc e−SY M F P . (55)

FT
where the action which must be considered for quantum theory is
SY M F P = SY M + Sgf , (56)
with
Z h  α A i
Sgf = d4 x bA ∂µ AA
µ + b + cA ∂µ DµAB cB . (57)
2
A few observations are in order:
RA
• The classical field equation of b results in the gauge fixing condition for f a = −αba (Exercise 1.29). Therefore,
with the help of ba , it is more evident that the gauge fixed action actually enforces the proposed constraint;
• The cases α = 0 and α = 1 are recognized as the Landau and Feynman gauges, respectively;
• The algebra-valued fields c = cA T A and c = cA T A are the Faddeev-Popov ghosts and anti-ghost fields. They
are fermionic fields with no Lorentz indexes (scalars), i.e., they are spin-0 bosons obeying Fermi statistics
(because of the Faddeev-Popov determinant). In principle, they should violate the spin-statistics theorem [59]
and, consequently, a causality violation is expected. However, what happens is that these fields are actually
non-physical because they do not appear as asymptotic states [6], i.e., they are not observables of the theory.
Nevertheless, together with the gauge fixing, they play an important role in canceling non-physical degrees of
freedom of the gauge fields, ensuring the unitarity of the theory. In fact, there is a class of gauge fixings where
the Faddeev-Popov fields decouple, the so called axial gauge nµ Aµ = 0, where nµ is a fixed vector [4];
D

• An important consequence of (56) is that it displays the Faddeev-Popov discrete symmetry, which means that
SY M F P is invariant by c 7−→ c and c 7−→ −c. This symmetry characterizes the ghost number, see Table I. Since
it is a symmetry of the action, all observables should also be invariant under the Faddeev-Popov symmetry,
which means that this symmetry prevents ghost states to appear in external legs of any diagram - so, causality
is safe indeed;
• In the case where G is Abelian, such as in QED, the Faddeev-Popov determinant can be eliminated by simple
normalization of Z (Exercise 1.30). Thus, QED is naturally ghost-free.
• The Yang-Mills action (56) has no mass parameters, i.e. the theory is massless and conformal at classical level.
• It is important at this point to expend a few words about the mass dimensions of the fields in four-dimensional
spacetime. Because ~ = 1, the action should be dimensionless in natural units. Thus, since [x] = −1 in mass
dimension, [ d4 x] = −4 and [∂] = 1. As a consequence, [L] = 4. A simple algebraic analysis leads to the second
R
column of Table I (Exercise 1.31);

13 In a way, the arbitrariness of f is replaced by α.


15

Fields A b c c̄ κ
Dimension 1 2 0 2 0
Ghost number 0 0 1 −1 0

TABLE I: Quantum numbers of the Yang-Mills fields in four-dimensional spacetime.

B. BRST quantization

A remarkable feature of the gauge fixed action (56) is that, in despite of the broken gauge symmetry, a new symmetry
shows up, the so called BRST symmetry [6, 43, 60, 61], due to C. Becchi, A. Rouet, R. Stora and, independently,
I. V. Tyutin. BRST symmetry is characterized by a fermionic operator s, acting on the fields as

sAA AB B
µ = −Dµ c , (58a)
κ
scA = f ABC cB cC , (58b)
2
scA = bA , (58c)

FT
A
sb = 0 . (58d)
It is a straightforward exercise to show that (Exercise 1.32)
sSY M GF = 0 . (59)
From (58a - 58d) it is clear that BRST symmetry is actually a supersymmetry because it transforms bosons in fermions
and vice versa. Moreover, the BRST operator is nilpotent, s2 = 0 (Exercise14 1.33), a very welcome feature. From
(58a - 58d) and Table I we also have that [s] = 0 and that s has ghost number 1. Hence, when acting over an object,
the BRST operator increases its ghost number by 1.
It turns out that this very beautiful symmetry plays a fundamental role in non-Abelian gauge theories. As briefly
RA
discussed in the Appendix B, s has a deep geometrical meaning, it is the exterior derivative in the functional space
of all gauge configurations along a gauge orbit. Moreover, the Faddeev-Popov ghost field is a Maurer-Cartan form
in this very same space while the gauge field is a connection over a more primitive fiber bundle structure [2, 62, 63],
which we have called gauge space. In fact, the Faddeev-Popov determinant corresponds to a Riemannian metric in
the gauge functional space. Thus, independently of the Physics (independently of the chosen action), A, c and s exist
as fundamental mathematical abstract structures.
This fundamental appeal is the origin of the so called BRST quantization method where the BRST symmetry is
taken as the fundamental principle to construct the invariant action. The main ingredients are the gauge field A, the
ghost field c and the BRST operator s. The anti-ghost c and the Lautrup-Nakanishi b fields are introduced in order
to account for the gauge fixing and also to ensure the Faddeev-Popov discrete symmetry. Before we proceed, let us
enunciate two important results about the BRST operator:
D

• The fact that s is a nilpotent operators allows one to define its cohomology, the BRST cohomology. The BRST
cohomology problem is the problem of finding the solution of15
sΘgn = 0 , (60)
where Θgn is typically and integrated object with specific ghost number g and dimension n. The solution of (60)
is [43]

Θn = Υn + sΘn sΥgn = 0 and Υgn 6= s(something) .
g g g−1

(61)

In (61), Υgn is the non-trivial part of the cohomology and sΘg−1


n is the trivial sector of the cohomology16 .

14 Tip: Use the Jacobi identities (21).


15 This is actually the fundamental cohomology problem for any nilpotent operator.
16 It is commonly used the terminology BRST exact object for trivial objects and BRST closed object for non-trivial objects
s(exact) = 0 ⇒ exact = s(something) ,
s(closed) = 0 ⇒ closed 6= s(something) . (62)
16

• BRST doublets consist on pair of fields transforming through sX = Y and sY = 0. Clearly, c and b are BRST
doublets. A general result is that BRST doublets belong exclusively to the trivial sector of the cohomology
[43]. In practice, any BRST doublet can be introduced because they are harmless to the physical content of the
theory.
Thus, to find the BRST invariant action it is imposed that S is a local polynomial in the fields and their derivatives
(with no higher order derivatives), with dimension bounded by 4, vanishing ghost number, Euclidean spacetime
isometry, obeying
sS = 0 . (63)
Condition (63) is actually the BRST equivalent of the gauge principle. Therefore, following (63), the solution is then
given by

−1

S = S0 + s∆ sS0 = 0 , S0 6= s(something) . (64)

where ∆−1 is a local integrated polynomial in the fields and their derivatives, with dimension bounded by 4 and
ghost number −1. The fact that BRST doublets always belong to the trivial sector combined to the Faddeev-Popov

FT
symmetry yield that c, b and c are not allowed at S0 , i.e. S0 = S0 (A). Hence, it can be shown that (Exercise17 1.34)
(see [43])
S0 = SY M ,
Z  α A 
∆−1 = d4 x cA ∂µ AA
µ + b + βf ABC B C
c c . (65)
2
where α and β are dimensionless gauge parameters. For the particular case β = 0, (64) reduces to the Faddeev-Popov
result(56) (Exercise 1.35). In fact, (57) can be written as a BRST exact object
RA
Z  α A
Sgf = s d4 x cA ∂µ AA
µ + b . (66)
2
which means that the gauge fixing does not interfere with the physical observables.
An example of a non-physical quantity is the gauge field propagator (Exercise 1.36),
1
hAA B
µ (−p)Aµ (p)i = δ
AB
[Tµν + αLµν ] , (67)
p2
where the transverse and longitudinal projectors read
pµ pν
Tµν = δµν − ,
p2
D

pµ pν
Lµν = . (68)
p2
The gauge field propagator is then clearly gauge dependent and describes massless non-physical states.

C. Renormalization, physical states and unitarity

There are three very important consequences of BRST symmetry: renormalizability, definition of physical states
and unitarity. Renormalizability states that all divergences can be consistently eliminated by a suitable redefinition of
the fields and parameters. This can be performed through the algebraic renormalization techniques [43]. Essentially,
BRST symmetry is imposed for the quantum action Γ, defined through
e−Γ = Z . (69)

17 This is not a trivial exercise. For the non-trivial part, write down a linear combination of all possible integrated polynomial terms using
A, ∂, κ, and f ABC . Impose the requirements above listed and then apply s. The result must vanish, a property that will provide
specific values for the coefficients of the linear combination. The final result should be the Yang-Mills action. The trivial part is a bit
more complicated due to the higher number of fields (which includes A as well).
17

The quantum action can be perturbatively expanded around the classical action,

Γ = Γ0 + Γ1 + 2 Γ2 + . . . ... (70)

where Γ0 = SY M F P ,  is a small expansion parameter and Γn , for n > 0, are integrated local polynomials in the
fields and their derivatives bounded by dimension four, vanishing ghost number and Euclidean Lorentz invariance.
Imposing18 sΓ = 0 and using sΓ0 = 0, it is possible to show that, at first order, Γ1 can be reabsorbed in Γ0 by means
of multiplicative redefinition of the fields and parameters, ensuring that Γ(1) = Γ0 + Γ1 has the same form as Γ0 .
The method is recursive and can be applied to all orders. Thus, Yang-Mills theories are renormalizable to all orders
in perturbation theory, at least in a large class of covariant gauges.
The gauge principle is used to determine physical observables at the classical level. In non-Abelian theories, the
BRST symmetry must be taken in favor of gauge symmetry because of the presence of the Faddeev-Popov ghosts.
Thus, (56) can also be considered at classical level. Therefore, since the BRST symmetry is postulated to be the
fundamental symmetry of Yang-Mills theories (it generalizes the gauge symmetry), it should be used to define the
physical states of the quantum theory. For that, it is used the nilpotent BRST charge Qbrst which is the generator of
BRST symmetry. From (Exercise19 1.36),
Z

FT
Qbrst = d3 j0brst , (71)

we can define the physical quantum states as BRST invariants,

Qbrst |ψi = 0 , (72)

where |ψi is required to have vanishing ghost number [64, 65] and to respect all symmetries displayed by SY M F P .
However, following the cohomological prescription (60) and (61), |ψi will have a non-trivial piece |ψiphys and a trivial
redundancy |ψinphys , namely
RA

Q |ψiphys = 0 |ψiphys 6= Qbrst |somethingi ,
brst


Q |ψinphys = 0 |ψinphys = Qbrst |somethingi .
brst

(73)

It is clear that |ψinphys has a vanishing norm (Exercise 1.38). Thus, the actual physical states are those belonging to
the non-trivial BRST cohomology in the space of vanishing ghost number states.
Unitarity of the S-matrix ensures the probability conservation in physical processes [6, 64, 65]. BRST symmetry
plays a fundamental role in that issue, guaranteeing that states with negative norm do not appear as asymptotic
states, i.e. they are not observables. The demonstration relies on the definition of physical states, Ward identities20
and commutation relations involving Qbrst .
D

D. Asymptotic freedom and confinement

The renormalization of the coupling parameter κ leads to the concept of asymptotic freedom [6, 66–68], which means
that the coupling parameter is higher as lower is the energy scale. In fact, at first order in perturbation theory, it is
found that [66, 67]

16π 2
κ2 = 2 , (74)
11N
3 ln Λµ2
qcd

18 All symmetries must be imposed as well.


19 Noether‘s theorem [6] establishes that
X ∂L
jµ = δΦA .
allf ields
∂ (∂µ ΦA )

20 Ward-Takahashi identities are functional equations describing the symmetries of an action. At quantum level, their role is to provide
recurrence relations between Green funtions.
18

where N is the Casimir of the gauge group (This is the dimension of the vector space of the fundamental represen-
tation), µ is the energy scale and Λqcd is the renormalization group cutoff [68]. The renormalization group cutoff is
a phenomelogical parameter which is associated with the pion mass. From (74) it is clear that as µ → ∞, κ → 0.
Thus, as higher the energy, more weakly coupled to each other are the gauge fields and perturbation theory is very
consistent. At this level, the gauge fields are almost free. On the other hand, as µ → Λqcd , µ → ∞. In this case, the
gauge fields become strongly coupled and perturbation theory does not hold anymore, i.e. the theory become highly
non-perturbative. The point µ = Λqcd is known as the Landau pole and indicates a phase transition of the theory
which is recognized as the hadronization phenomenon [69]. Below µ = Λqcd the Yang-Mills action must be replaced
by a suitable theory where the excitations are hadrons and glueballs21 .
The concept of confinement emerges from the behavior (74) together with the global G symmetry22 . The essence
of the global symmetry is to forbid any algebra-valued to be a physical state, Qbrst |ψ A >6= 0. Only states carrying
no group indexes would allowed at the physical sector. This property should hold at all energy scales23 . Even at
high energies, where the coupling is very small and the gauge excitations barely interact among each other, is also
impossible to observe them due to its algebra-valued character. At this regime, considering also quarks, the theory
describes the so-called quark-gluon plasma, a phase in which quarks and gluons are virtually free, even though it is
impossible to isolate a single quark or gluon. By decreasing the energy, the coupling increases and the interactions
between gluons (and quarks) become stronger and stronger til the point of hadronization at the Landau pole. Then,

FT
hadrons and glueballs, which are experimentally observables, are in fact states with no gauge charge.

III. GRIBOV AMBIGUITIES

In this section, we will introduce the problem pointed out by V. N. Gribov in his seminal paper [14]. The essence
of the problem lies on the fact that the Faddeev-Popov method, described in Sect. II, is not sufficient to fix the gauge
redundancy completely. It means that, after applying the Faddeev-Popov gauge fixing procedure, a residual symmetry
remains: there still are configurations obeying the same gauge fixing and related through gauge transofrmations. Such
spourious configurations are called Gribov copies and, although present in a wide class of gauges, their elimination
RA
is not fully understood. The first (partial) attempt to deal with such copies was realized by Gribov himself, and it
was worked out in more generality by Zwanziger later, [70]. A local and renormalizable action, free from infinitesimal
Gribov copies (copies that are related to another configuration through an infinitesimal gauge transformation) was
established, the so-called Gribov-Zwanziger action. This action was firstly constructed in the Landau gauge and
it has a very interesting feature: the term which is responsible to guarantee the elimination of copies breaks the
BRST symmetry (58a-58d) in a soft manner24 . Further non-perturbative effects as the presence of condensates can
be taken into account in this scenario giving rise to the Refined Gribov-Zwanziger action, [71], which provides gluon
and ghost propagators in very good agreement with lattice results. In this section, we provide an introduction to
these developments in a very concise form. For further details, we refer to more complete treatments which are listed
along the text.
D

A. Statement of the problem

As discussed in Sect. II, the quantization of gauge theories requires the imposition of a condition over the gauge
fields. Although the Faddeev-Popov framework (or BRST quantization) provides a very efficient way to fix a gauge
in perturbation theory, it fails at the non-perturbative level. The reason is the following: The application of the
FP procedure tacitly assumes that the gauge fixing condition has a unique solution for each gauge orbit (44). This
assumption does not hold when we enter the non-perturbative sector and spurious configurations are not eliminated
from the path integral quantization. As a matter of illustration, we consider pure Yang-Mills action in four-dimensional
Euclidean space with SU (N ) gauge group25 and we impose the Landau gauge condition over the gauge fields, which

21 This is actually an open problem in theoretical physics.


22 In QCD this symmetry is the color symmetry associated with the SU (3) global group. In electroweak interactions the SU (2) × U (1)
local and global symmetries are broken to the U (1) group due to spontaneous symmetry breaking. Hence, confinement is not a feature
of electroweak interactions.
23 There are no experimental evidence of confinement violation whatsoever.
24 A breaking is called soft if it is an operator with dimension lower than spacetime dimension. Otherwise, it is a hard breaking.
25 The group SU (N ) is chosen to fit with the existing literature. The generalization to a generic semi-simple Lie group G can be
straightforwardly done.
19

is the case of α = 0 in (57)

∂µ AA
µ = 0. (75)

An equivalent gauge configuration A0 is obtained from A through a gauge transformation, given by eq.(33). If we
assume that A satisfies the condition (75), the gauge fixing is consistent if eq. 75 is not satisfied by a configuration
A0 , which belongs to the same gauge orbit of A. To simplify the question, let us consider an infinitesimal gauge
transformation26 with gauge parameter ξ as in (34). Therefore,

A0A A AB B
µ = Aµ + D µ ξ , (76)

satisfies the Landau gauge condition if

∂µ A0A A AB B
µ = ∂µ Aµ + ∂µ Dµ ξ ⇒ ∂µ DµAB ξ B = 0 . (77)

It implies that a Gribov copy exists if the operator −∂µ Dµab , which is the Faddeev-Popov operator in the Landau
gauge, has zero-modes. Although this is not a trivial task, we can start our analysis as an eigenvalue problem, i.e.,

FT
MAB = −∂µ DµAB ξ B = −∂µ (δ AB ∂µ − κf ABC AC B A
µ )ξ = (A)ξ , (78)

where  is the correspondent eigenvalue associated to the gauge configuration A. This equation can be faced as a
Schrödinger-like equation, where A plays the role of the potential. An extra item to this analysis is that the operator
−∂D, with the condition ∂A = 0, is Hermitian (Exercise 1.39) and, therefore, has a real spectrum. From eq.(78), it
is possible to realize that for small values of A, just positive  are allowed due to the positivity of −∂ 2 in Euclidean
space. However, for a suitable value of A, as it becomes larger, the operator −∂D admits a zero-mode and, then, a
negative eigenvalue. This allows the following picture: Around A ≈ 0, i.e., at the perturbative level, the FP operator
does not have any zero-modes and the usual gauge-fixing procedure is safe. Nevertheless, as far as we get from the
perturbative sector, the FP operator reaches zero-modes and the gauge fixing is ill-defined. From this qualitative
RA
analysis, it is possible to see that the appearance of Gribov copies is a genuine non-perturbative effect.
This poses a non-trivial issue: The Faddeev-Popov gauge fixing procedure does not take into account the existence
of Gribov copies at the non-perturbative regime. Therefore, the path integral quantization is not free of spurious
configurations, which must be eliminated anyway. At the present moment, we have discussed just the existence of
infinitesimal copies. However, dealing with such class of copies is already highly non-trivial and very few is known
about the general problem with finite gauge transformations.
Before ending this general discussion, we must highlight a very important remark: In the way we presented the
existence of copies, might seem to the reader that this is a peculiarity of the Landau gauge. However, it was shown
by I. M. Singer [15] that the existence of such configurations is a very general feature, and, in particular, all Lorentz
covariant gauge conditions suffer from this problem. This enforces the reasoning that this problem is much more
delicate than a simple technical issue. For a more formal justification, see appendix B.
D

B. Gribov’s solution

In his seminal work, [14], Gribov proposed the first attempt to eliminate gauge copies. The idea, at a first glance,
might seem very naive and many drawbacks can be pointed out. However, many essential properties which were not
proved by Gribov in his paper were proved later on, providing a firm ground to his proposal. Therefore, although it
is not the final answer we have so far, the solution proposed by Gribov is simple and elegant enough to be presented
before the more recent attempts, namely, the Gribov-Zwanziger framework. We reenforce, however, that we will just
present the main ideas and we refer to appropriate references for more details. In particular, our exposition is based
on [45, 72, 73]. Furthermore, unless the contrary is said, we will restrict ourselves to the Landau gauge.
As mentioned before, the FP operator is Hermitian in the Landau gauge. Hence, its spectrum is real and, therefore,
it is possible to define positivitness and a hierarchy ordering for the eigenvalues . In particular, we argued that there
is a region where this operator is positive. So, inside this region, the operator does not contain zero-modes and no
infinitesimal Gribov copies are present27 . Gribov’s proposal was simply to restrict the domain of integration of the

26 We will only consider infinitesimal gauge transformations and, hence, only infinitesimal Gribov copies.
27 This does not prevent the appearance of finite Gribov copies.
20

path integral to this region. With this, we eliminate a large set of copies, namely, the infinitesimal ones. But, before
implementing such idea, we must highlight some important properties of this region.
The region Ω where MAB ≡ −∂µ DµAB is strictly positive is known as the first Gribov region, or simply Gribov
region. Therefore,

Ω = AA A AB

µ , ∂µ Aµ = 0, M >0 . (79)

This region enjoys some very important properties. In particular,


• it is bounded;
• it is convex;
• all gauge orbits cross Ω;
• the configuration AA
µ = 0 belongs to Ω.

The first two properties ensures that configurations very close to the so-called Gribov horizon, the boundary which
contains the first zero-modes of the FP operator, have a finite distance to the origin in configuration space. The third

FT
property is fundamental: It guarantees that the restriction is consistent, i.e., the restriction of the path integral to
Ω does not eliminate any physical configuration, which is essential. The proof of such properties can be found in
[74, 75]. Finally, the fourth property ensures that the perturbative region is inside the Gribov region and its proof is
trivial. These properties make the restriction a consistent proposal.
A very important comment is the following: The restriction to Ω, although satisfies those very important properties
does not eliminate all Gribov copies. As mentioned before, Gribov copies related by finite gauge transformations may
belong to Ω, [76]. The region which is truly free from copies is known as the fundamental modular region (FMR).
Nevertheless, we do not have a systematic way to implement the restriction of the path integral to this region so far.
So, the best we can do is to restrict to Ω.
Formally, for the restriction to Ω, we write for the path integral the following modified measure,
RA
Z
−SY M F P
Z = N DADc̄Dc δ(∂µ AA µ )e V(Ω) , (80)

where N stands for a possible normalization factor and V(Ω) is the factor responsible for the restriction to Ω.
Essentially, the factor V(Ω) is a Heaviside function limiting the functional integral up to the Gribov horizon. To
characterize V(Ω) we will first notice that, in the Landau gauge, the operator which generates infinitesimal Gribov
copies, namely, the FP operator, is intimately related to the ghost sector of the theory. In particular,
Z
 −1 AB
hc̄A (x)cB (y)i = DAδ(∂µ AA µ )V(Ω)det(M) M (x, y)e−SYM (81)

where we have integrated over the ghost fields. Thus, performing the one-loop computation of the ghost two-point
D

function within the usual Faddeev-Popov theory (55), we obtain the following result [14, 45] (Exercise 1.40)
1 1
hc̄A (−p)cB (p)i = δ AB G(p) = δ AB 9 , (82)
p2  11κ2 N Λ2
 44
1− 48π 2 ln p2

 2

with Λ being the ultraviolet cutoff. The function G(p) has two singularities: p2 = 0 and p2 = Λ2 − κ12 48π
11N . For
2
large momenta, G(p) ≈ 1/(p2 ln Λp2 ) which is the perturbative regime and we are inside the Gribov region as we
 2

discussed before. Also, it is possible to note that for p2 < Λ2 exp − κ12 48π
11N , G becomes complex, which means that
we are outside the Gribov region28 Ω. On the other hand, the pole p2 = 0 corresponds to the fact we have reached
the Gribov horizon, since p2 is a positive function. This implies that the restriction to the Gribov region avoids
singularities different from p2 = 0. In fact, this is the away we can impose the restriction of the path integral, i.e.,
let us impose that no poles but p2 = 0 are allowed in the ghost two-point function. This is the so-called no-pole
condition, which was the original proposal by Gribov and allows the characterization of V(Ω). We must highlight that

28 A similar discussion will take place at Sec. X C


21

this proposal was worked out up to second order in perturbation theory originally. The generalization to all orders
will be discussed in the next section.
For the time being, we compute the connected ghost two-point function up to second order using the usual FP path
integral (55). We will omit the details, but they can be found in [45, 72, 73]. The idea of using the (55) is because
we want to find a condition for the ghost propagator in the usual theory that identifies the Gribov horizon, i.e. we
want to identify the FP zero modes with the usual ghost propagator. Then, this identification will be used to define
V(Ω). Moreover, A must be considered as external fields in such a way that V(Ω) will be a function of A itself. The
G function is then

1 1 N κ2 d4 q A (p − q)µ pν
Z
1
G(p, A) = 2 + A (−q)AA ν (q) , (83)
p V p4 N 2 − 1 (2π)4 µ (p − q)2

and V is the spacetime volume. We rewrite (83) as


1
G(p, A) = (1 + σ(p, A)) , (84)
p2
with

FT
1 1 N κ2 d4 q A (p − q)µ pν
Z
σ(p, A) = A (−q)AA
ν (q) . (85)
V p2 N 2 − 1 (2π)4 µ (p − q)2

Then, (84) is rewritten in Born approximation as


1 1
G(p, A) ≈ 2
, (86)
k 1 − σ(p, A)

and the no-pole condition can be simply implemented by σ(p, A) < 1. Although we will not give the proof [72, 73],
σ(p, A) decreases as p2 increases. So, the no-pole condition can be rephrased as σ(0, A) < 1 where [45] (Exercise 1.41)
RA
1 1 N κ2 d4 q A
Z
1
σ(0, A) = 2 4
Aκ (−q)AA
κ (q) 2 . (87)
V 4N −1 (2π) q

Hence, an explicit expression for V(Ω), in terms of σ, can be written:

V(Ω) = θ(1 − σ(0, A)) . (88)

Even though we have determined an expression for V(Ω), in its current form expression (88) is not suitable for
practical purposes. Nevertheless, we can incorporate V(Ω) into the classical action using an integral representation
for the Heaviside function, i.e.
D

Z +i∞+
dβ β(1−σ(0,A))
V(Ω) = θ(1 − σ(0, A)) = e . (89)
−i∞+ 2πiβ

Thus,
Z +i∞+ Z
dβ −SY M F P β−ln β−βσ(0,A))
Z=N DADc̄Dc δ(∂µ AA
µ )e e , (90)
−i∞+ 2πi

The trick here is to set β to its minimizing value (the saddle point method [6, 14]), i.e. δZ/δβ = 0, providing the so
called gap equation (Exercise 1.42),

d4 q
Z
3 2 1
1 = Nκ , (91)
4 (2π)4 (q 4 + γ 4 )

where the so called Gribov parameter γ is defined as


β0 N 1
γ 4 = κ2 , (92)
N 2 − 1 2V
22

and β0 is the solution of the saddle point. Two observations are in order: First, the actual value of γ is the one
determined by (91). Second, in (92), the rate β0 /V is assumed to be finite as V 7−→ ∞. Hence, after a suitable
normalization, we finally have29 (Exercise 1.43)
Z R 4 2 4 4
−SY M F P
Z = DADc̄Dc δ(∂µ AA µ )e e d x4(N −1)γ −γ σ(0,A)) , (93)

where the thermodynamic limit V 7−→ ∞ was taken. the term γ 4 σ is clearly momentum-dependent and, thus, a
non-local term while the V γ 4 is a vacuum term30 . However, since the new term is quadratic in the gauge field, it
affects the tree level propagator which is now given by (Exercise 1.44)

p2
hAA B
µ (p)Aν (−p)i = δ
AB
Tµν . (94)
p4 + γ4

where, as we have established, the Landau gauge is assumed (α = 0).


Some comments are (focusing on QCD), however, important: The restriction of the path integral to the Gribov
region brings deep modifications to the gluon propagator. In particular, it is suppressed in the infrared region and
displays two complex poles. This makes the interpretation of gluons as part of the physical spectrum impossible,

FT
which can be faced as a manifestation of confinement. It is a quite remarkable feature that the partial solution of a
technical problem provides such physical behavior. The computation of the ghost propagator also has novelties, but
we will omit it here because it is not relevant to our present purposes: its behavior is enhanced in the infrared regime
at one-loop order. Its behavior is ∼ 1/k 4 . Finally, the integral in the gap equation (91) is divergent. Thus, to solve
it, renormalization is required. A local and renormalizable action which effectively implements the restriction to the
Gribov region Ω is then needed. This is achieved by the so-called Gribov-Zwanziger action, which will be exposed in
the next section.

C. The Gribov-Zwanziger action and BRST soft breaking


RA
In the last section, we discussed the solution proposed by Gribov for the elimination of infinitesimal gauge copies.
Nevertheless, this solution was implemented up to second order in perturbation theory and we might wonder if an
all order proposal could be developed. The answer is affirmative and was first worked out by D. Zwanziger in [70].
Inhere, again,we will not expose all the reasoning followed by Zwanziger, but just mention some important features.
Instead of dealing with an all order no-pole condition, Zwanziger studied the spectrum of the Faddeev-Popov
operator in the Landau gauge, which is actually the same thing [77, 78]. In particular, the restriction to the Gribov
region is implemented through the imposition that the functional integral should be performed in the region of
configuration space where the smallest eigenvalue of M = MAA /(N 2 − 1), i.e. (A), is non-negative. We must notice
that the eigenvalue is a function of the gauge field. Zwanziger argued that this condition is well implemented, in the
thermodynamic limit, by requiring the trace of M is non-negative. The condition found by Zwanziger is given by
D

4V (N 2 − 1) − H(A) ≥ 0 , (95)

with
Z
 −1 AD
H(A) = g 2 d4 xd4 y f ABC AB
µ (x) M (x, y)f DEC AE
µ (y) , (96)

being the so-called horizon function. Also, Zwanziger was able to effectively implement (95) by a modification in
the classical action. The result for the modified path integral which takes into account the restriction to the Gribov
region Ω is given by
Z
−[SYM +γ 4 H(A)− d4 x4(N 2 −1)γ 4 ]
R
Z = DAδ(∂µ AA µ )det(M) e , (97)

29 An important step here is that the term ln β can be neglected at the thermodynamic limit because the Heaviside function, at this limit,
converges to the Dirac delta function. See [45].
30 This term is negligible at classical level. At quantum level, however, it is essential to obtain the gap equation (91)
23

where γ - the Gribov parameter - is not free, but determined in a self-consistent way through the so-called horizon
condition,

hH(A)iGZ = 4V (N 2 − 1) , (98)

with the expectation value being evaluated with the path integral modified measure. It is possible to show that
eq. (98) reduces to the gap equation given by (91) at lowest order, i.e., in the Gribov approximation.
From (97), it is possible to read the modified classical action, known as the Gribov-Zwanziger action, which imple-
ments the restriction to the Gribov region Ω (to all order no-pole condition), i.e.
Z Z
1 4 A A
SGZ = d x Fµν Fµν + d4 x (c̄A ∂µ DµAB cB + bA ∂µ AA 4 4 2
µ ) + γ H(A) − 4γ V (N − 1) , (99)
4

which, due to the presence of H(A), is a non-local action. Nevertheless, it can be casted in a local form through the
introduction of a set of auxiliary fields, namely, a pair of commuting (ϕ̄AB AB
µ , ϕµ ) and a pair of anti-commuting fields
(ω̄µAB , ωµAB ). With this new set of fields, the Gribov-Zwanziger action is expressed as (Exercise31 1.45)

FT
Z Z Z
1
SGZ = d x Fµν Fµν + d x (c̄ ∂µ Dµ c + b ∂µ Aµ ) − d4 x ϕ̄AC
4 A A 4 A AB B A A
µ M
AB BC
ϕµ − ω̄µAC MAB ωµBC
4
Z
BC
− gf ABC ∂µ ω̄νAD DµBE cE EϕCD 2
d4 x f ABC AA 4 2

ν − gγ µ (ϕ̄ + ϕ)µ − 4V γ (N − 1) . (100)

By integrating over the auxiliary fields, we regain expression (99) (Exercise 1.46). The auxiliary fields, also known as
Zwanziger fields, form a BRST quartet (two BRST doublets),

sϕAB
µ = ωµAB , sωµAB = 0 ,
sω̄µAB = ϕ̄AB
µ , sϕ̄AB = 0. (101)
RA
µ

As a first test, it can be shown that the action (101) reduces to the usual gauge fixed Yang-Mills action in the limit
γ → 0 (Exercise 1.47).
The Gribov-Zwanziger action has a very peculiar property: It breaks explicitly the BRST symmetry (Exercise
1.48). Albeit explicit, the breaking is soft because it is proportional to the Gribov parameter. It means that, in the
ultra-violet regime, the Gribov parameter becomes negligible and the BRST symmetry is recovered (together with all
well established properties of perturbative quantum theory). Hence, the explicit breaking of the BRST symmetry is
a direct consequence of the restriction of the domain of the path integral to the Gribov region. Since the restriction
to the Gribov region allows the interpretation of confinement through the modification of the gluon propagator, we
immediately realize that there is a deep connection between confinement and BRST soft breaking.
It is important to mention that the Gribov-Zwanziger action is renormalizable to all orders in perturbation theory.
D

The proof is beyond the scope of these notes, but we refer to [70, 72, 73] for detail.

D. The Refined Gribov-Zwanziger action

Besides the restriction to the Gribov region, which is a non-perturbative effect in the sense that in the perturbative
regime Gribov copies play no role, other non-perturbative features might be taken into account. Doing so, a suitable
modification of the Gribov-Zwanziger framework is achieved and its final form fits very well with the most recent
lattice data [79–81]. The action which leads to those results is the so-called Refined Gribov-Zwanziger action [71].
Essentially, the Gribov-Zwanziger action displays non-vanishing values for condensates, in particular, for hAaµ Aaµ i and
hϕ̄ab ab ab ab
µ ϕµ − ω̄µ ωµ i. The non-vanishing values at one-loop are proportional to the Gribov paramete [71, 82]. Therefore,
the restriction to the Gribov region brings more non-trivial features at the non-perturbative level. The proposal is
to take into account these instabilities of the Gribov-Zwanziger action from the very beginning by introducing those

31 The demonstration is performed in the full path integral (97).


24

condensates. It implies a modification to the Gribov-Zwanziger action as follows


Z Z Z
1
SRGZ = − d4 x Fµν
A A
Fµν + d4 x (c̄A ∂µ DµAB cB + bA ∂µ AA µ ) − d4 x ϕ̄AC
µ M
AB BC
ϕµ − ω̄µAC MAB ωµBC ,
4
m2
Z Z
BC
− κf ABC ∂µ ω̄νAD DµBE cE ϕCD 2 4 ABC A
d4 x AA A

ν − κγ d x f A µ (ϕ̄ + ϕ) µ + µ Aµ + ,
2
Z Z
2
−M d x (ϕ̄µ ϕµ − ω̄µ ωµ ) − d4 x4γ 4 (N 2 − 1) ,
4 AB AB AB AB

(102a)

which is the Refined Gribov-Zwanziger (RGZ) action. This action, as the Gribov-Zwanziger action, breaks the BRST

FT
symmetry in a soft manner. The gluon propagator computed from the RGZ action is (Exercise 1.49)

p2 + M 2
hAA B
µ (p)Aν (−p)i = δ
AB
Tµν , (103)
(p2 + m2 )(p2+ M 2 ) + 2g 2 N γ 4

which, as previously discussed, does not vanish at the origin p → 0. An important remark is that for m2 = 0, it still
goes to a non-vanishing value, due to the presence of the condensates of auxiliary fields which naturally emerge from
the localization of the horizon function. The ghost propagator is not enhanced anymore in this framework, i.e., as
G(p2 ) ≈ 1/p2 .
Some considerations are in order: The Refined Gribov-Zwanziger scenario has passed through several non-trivial
RA
tests in the last few years. In particular, besides the very good agreement with lattice results for the gluon and ghost
propagators, [79–81], it also provides a good qualitative agreement with glueball spectrum [83, 84], the correct sign
to the Casimir energy in the MIT bag model [85] and a coupling to scalar and fermionic matter that also fits lattice
simulations [86]. As presented here, the (Refined) Gribov-Zwanziger action might seem to be restricted to the Landau
gauge. Indeed, its construction is based on very peculiarities of Landau gauge and it was the first gauge condition
which was worked in details. However, we must mention that we already have at our disposal the extension (or,
at least, the partial extension) to other gauges, e.g., the maximal abelian, linear covariant and Landau - maximal
Abelian interpolating gauges, see [49, 82, 87, 88]. Also, many aspects of the soft BRST breaking are being studied
not only from the analytical point of view, but also from lattice simulations [50, 89]. This scenario seems to be a
very suitable way to study non-trivial effects from the non-perturbative regime of Yang-Mills theories in an analytical
fashion.
The issue of the BRST symmetry breaking is highly non-trivial and a matter of current research, its final status
D

still lacks. It is worth mention, however, that a non-perturbative version of the BRST operator has been found in
[50] where its action on the RGZ fields are highly non-local. Nevertheless, this new BRST operator is nilpotent and
reduces to the ordinary BRST operator at the perturbative level.
25

PART II: GRAVITY AS A GAUGE THEORY

In this part we focus on the construction of gravity in the first order formalism. Some definitions will be made in
order to clarify the main differences between first and second order formalisms, also known as vierbein and metric
formalisms, respectively. The main references on general relativity in the first and second order formalisms are
[10, 12, 90–98].

IV. EQUIVALENCE PRINCIPLE AND GEOMETRODYNAMICS

Following Einstein’s ideas on the construction of general relativity (GR) [99], gravitational effects can be locally
eliminated for an object in free falling. By “locally”, it is meant a small spacetime region where gravity does not vary
significantly. A famous example is the free falling elevator experiment. This property is the essence of the modern
interpretation of the equivalence principle. Let us quickly discuss two equivalence principle enunciates:
• Inertial and gravitational mass are equivalent.
This enunciate is known as Weak Equivalence Principle (WEP), which states that the inertial mass of a body

FT
is equal to its the gravitational mass32 . The consequence is that all test particles locally experiment the same
acceleration under gravity. This principle is known since Galileo Galilei.
• For local spacetime regions, special relativity laws hold.
This version is known as the Strong Equivalence Principle (SEP). It explicitly states that gravitational fields
cannot be detected by local experiments. However, for sufficient large spacetime regions, gravitational field
inhomogeneities cannot be neglected since tidal forces take place.
The WEP states that gravity effects can be locally eliminated by a suitable non-inertial reference frame. The
SEP aditionally states that, locally, spacetime is a Minkowski flat spacetime. Hence, it is intuitive to try to describe
gravitational effects as a dynamical geometric formulation of spacetime which, locally, is reduced to spetial relativity.
RA
In this formulation, spacetime is deformed by matter while matter experiments a deformed spacetime due to the
presence of other matter distributions. Thus, the spacetime geometry is dynamically generated by the presence of
matter. This type of theories are called geometrodynamical theories or, simply, geometrodynamics.
These ideas lead Einstein to formulate a special set of equations describing spacetime dynamics due to matter
distributions, the so called Einstein equations,
1
R − 2Λ2 gµν = kTµν .

Rµν − (104)
2
where33 Rµν is the Ricci tensor (related to spacetime curvature), R = Rµν g µν is the curvature scalar, gµν is the
metric tensor, Λ is the cosmological constant, k is related to Newton’s constant (it can be determined by taking the
Newtonian limit of (104)), and Tµν is the energy-momentum tensor of the matter distribution. The l.h.s of (104) is
D

purely geometric34 while the r.h.s is related to pure physical properties of the matter distribution. Essentially, given
a matter distribution, (104) determines the geometric properties of spacetime. In this scenario, the metric tensor is
the fundamental field.
It is worth mention that the trace of (104) leads to (Exercise 2.1) R = 4Λ2 − kT where T = T µµ . Thus, even in the
absence of matter distributions, a curvature can be generated by the cosmological constant, R = 4Λ2 , which is known
to exist with an extremely small observational value (∼ 10−44 peV 2 ). In fact, the cosmological constant in (104) is
recognized as the dark energy component of the Universe.

A. General relativity in the metric formalism

The metric formalism of GR, or second order formalism, is based on the principle that gµν is the only fundamental
field of gravity. This is actually the original idea adopted by A. Einstein in the construction of (104). All other
geometric quantities are obtained from the metric tensor.

32 The gravitational mass is the equivalent of the electric charge in Coulomb force, which cannot be associated to inertia.
33 We will dissect the geometric quantities of a differential manifold in the next section.
34 Perhaps Λ can have a different origin other than geometry.
26

Let us start our definitions with the metric tensor gµν , which is the fundamental field of GR. Inherent to this
geometrical object are all information about lengths, angles, areas, and volumes on a differentiable manifold M 4 . It is
responsible to describe and to characterize the spacetime metric properties. For simplicity we will omit the position
dependence of most quantities. With gµν , the distance between two near points, xµ and xµ + dxµ , is easily computed,

ds2 = gµν dxµ dxν . (105)

Clearly, the metric tensor is a symmetric second order tensor, i.e., gµν = gνµ . An inverse metric tensor is assumed,
g µν , through

gνλ g λµ = δνµ , (106)

The SEP states that, locally, gµν ≈ ηµν , where ηµν is the Minkowski metric tensor. In the absence of a gravitational
field, ηµν becomes global in the spacetime.
The first object we can construct from the metric tensor is the affine connection, which characterizes the parallelism
properties of spacetime. The affine connection rises from the study of parallel transport in M 4 of a generic vector
field. It can also be obtained from the imposition that a covariant derivative of a vector must transform as a second
rank tensor. Hence (Exercise 2.2), lit can be defined the action of the covariant derivative as

FT
∇β T µ1 µ2 ···µkν1 ν2 ···ν` = ∂β T µ1 µ2 ···µkν1 ν2 ···ν`
+ Γµβλ1 T λµ2 ···µk ν1 ν2 ···ν` + Γµβλ2 T µ1 λ···µk ν1 ν2 ···ν` + ···
− Γλβν1 T µ1 µ2 ···µkλν2 ···ν` − Γλβν2 T µ1 µ2 ···µkν1 λ···ν` − ··· . (107)

where the affine connection Γµβν is defined under two strong requirements:
• Metric compatible: The metric tensor is covarantly constant, ∇β gµν = 0.
• Torsion free: T αµν = Γα α α α
µν − Γνµ = 0, which implies Γµν = Γνµ .
RA
In general, the affine connection is independent of the metric. In the metric formalism, however, it is only a function
of gµν . These requirements state that non-metricity and torsion vanishes at all spacetime points. Thus, parallel
transport preserves lengths and angles and also forbids twists in M 4 . Under these conditions, the affine connection is
reduced to the so called Christoffel symbols (Exercise 2.3)
1 αρ
Γα
µν = g (−∂ρ gµν + ∂µ gρν + ∂ν gµρ ) . (108)
2
This is the connection adopted by Einstein in the original formulation of the GR.
The Riemann-Christoffel tensor, or simply Riemann tensor, corresponds to the spacetime curvature. This tensor is
defined through the commutation of covariant derivatives over a vector, i.e. (Exercise 2.4)
D

[∇µ , ∇ν ] V α = Rαβµν V β , (109)

where

Rαβµν = ∂µ Γα α α λ α λ
νβ − ∂ν Γµβ + Γµλ Γνβ − Γνλ Γµβ . (110)

The main properties of the Riemann tensor in the metric formalism are
• It is antisymmetric in the first and last pair of indexes: Rαβµν = −Rαβνµ and Rαβµν = −Rβανµ .
• It is symmetric by interchanging the first and second pairs of indexes: Rαβµν = Rµναβ .
• Because the covariant derivative defines parallel transport, the curvature measures the failure of the parallelo-
gram law. In other words, parallel transport of a vector along two different directions will differ by interchanging
the order of the directions.
The Riemann tensor can be used to calculate two important objects present at Einstein equations (104). These
quantities are the Ricci tensor and the curvature scalar, namely, Rµν = g αβ Rβµαν = Rαµαν and R = g µν Rµν ,
respectively. We notice however that, ultimately, the Riemann tensor depends exclusively on the metric tensor.
27

It is possible to obtain the Einstein equations (104) from a variational principle. This is achieved by defining the
Einstein-Hilbert action,
 

Z
4 1
SEH = d x g (R − 2Λ) + T µµ , (111)
16πG

where g = | det gµν | and G is the universal gravitational constant (Newton’s constant). It is a straightforward
computation to obtain (104) from (111) (Exercise 2.5).
It is worth mention that (104) encondes Newtonian gravity. In fact, by taking the weak field limit we recover the
Newton’s theory of gravitation where g00 is related to the Newtonian potential. In this way, gµν would play the role
of a relativistic potential while Γα
µν would be the associated gravitational force. This interpretation is also evident
from the geodesic equation

d2 xµ α
µ dx dx
β
= −Γ αβ . (112)
ds2 ds ds
where s is the line element of the geodesic and xµ the position of the particle along the geodesic curve. This

FT
interpretation is consistent. However, an alternative interpretation can be defined in order to make gravity closer to
gauge theories. This is the task of the next section.

B. First order formalism of gravity

The most direct generalization of the metric formalism is due to A. Palatini and E. Cartan [100–102], which considers
gµν and Γµαβ as independent dynamical variables, i.e. there are two independent fields for gravity. In this way, new
degrees of freedom rises. This formalism is called Palatini formalism and brings torsion into the game. The effect is a
new set of field equations δSEH /δΓ = 0. Nevertheless, no torsion shows up unless the energy-momentum tensor of the
RA
matter distribution carries spin-density components (Exercise 2.6). Nevertheless, non-metricity remains a vanishing
quantity. Theories which nonmetricity is included are usually called metric-affine gravity [32].
The so called first order formalism [94] is equivalent to the Palatini formalism, but it is written in terms of different
variables, namely, the vierbein 35 eaµ (x) and the spin-connection ωµab (x) where small Latin indexes refer to the M 4
tangent space coordinates. In particular, for economical purposes, we will now employ the differential form notation36
and use the vierbein one-form ea (x) and the spin-connection one-form ω ab (x). Let us quickly construct these objects.

1. Vierbein

Let spacetime be a smooth37 four-dimensional manifold M 4 . At each point x ∈ M 4 we define the tangent space
Tx (M 4 ) with flat metric given by the Minkowski metric, η = diag(−1, 1, 1, 1). Due to the SEP, tangent spaces of
D

this form are equivalent to free falling local reference frames. In this way, we start a description of spacetime in full
explicit accordance with the SEP. By definition of tangent spaces, there is an isomorphism between Tx M 4 and M 4 ,
which is a linear map relating coordinates y of the tangent space and coordinates x of M 4 . This map is local, is called
vierbein, and is given by

dy a = eaµ dxµ , (113)

where, due to dim M = dim T (M ), we also have a, b, c . . . ∈ {0, 1, 2, 3}. Sometimes, the tangent indexes are called
frame indexes while spacetime indexes are called world indexes. Because a distance measure must be invariant under
this isomorphism, we must have the following relation (Exercise 2.7)

gµν ≡ ηab eaµ ebν . (114)

35 In d dimensions eaµ is called vielbein and for specific dimensions we have zweibein, dreibein, vierbein... for 2, 3, 4... dimensions.
36 A brief exposition about differential forms can be found in Appendix (A). For detailed texts we refer to []. In our notation, all wedge
products are implicit understood, and, hence, will be omitted in most equations.
37 By smoth, it is meant a C ∞ class manifold, i.e. an infinitely differentiable manifold.
28

Thus, a metric gµν on M 4 is induced by the vierbein eaµ and the tangent space metric ηab . The inverse vierbein is
defined through the set of relations:

eaµ eµb = δba ,


eaµ eνa = δµν . (115)

The flat Minkowski metric is characterized by the isometries of the tangent space. These isometries are described by
the Lorentz group SO(1, 3), but the local version of the Lorentz group, i.e. Lorentz transformations which depend on
the spacetime position. Thus, eaµ is not unique because there are infinity equivalent virbeins related by the elements of
the local group SO(1, 3). In other words, locally, there are infinity equivalent tangent spaces related SO(1, 3). Under
local Lorentz transformations, the vierbein behaves as

eaµ 7−→ Lab ebµ , (116)

where L = L(x) ∈ SO(1, 3). It is worth mention that the vierbein has n2 = 16 independent components whereas
the metric has n(n + 1)/2 = 10 independent ones. The difference is exactly the n(n − 1)/2 = 6 independent n = 4-
dimensional local rotations characterized by the Lorentz group.

FT
Ultimately, we define the vierbein one-form, which refers to a soldering form in the cotangent space Tx∗ (M 4 ), which
is dual to Tx (M 4 ), by

ea = eaµ dxµ , (117)

where dxµ is a local basis of Tx∗ (M 4 ). Clearly, (118) turns to

ea 7−→ Lab eb , (118)

which, for infinitesimal transformations reduce to (Exercise 2.8)


RA
ea 7−→ ea + ξ ab eb , (119)

where ξ ab is the infinitesimal group parameter. Obviously, due to its algebra-valued character, ξ ab = −ξb a .
An important observation is that the action of the Lorentz group on the vierbein leaves unaltered the spacetime
coordinates xµ but affects the vierbein. Thus, the local Lorentz group can be visualized as a gauge group for gravity.
Under this assumption, through (114), the metric tensor is a gauge invariant (Exercise 2.9). On the other hand, (118)
does not describe the gauge transformation of a connection but of a matter field! Thus, it remains to determine what
quantity is the connection in this scenario.
D

2. Spin-connection

Once again, the concept of parallel transport concept can be used to define the connection. Inhere, we are interested
in tangent and cotangent spaces. Let x and x + dx two near points on the spacetime manifold M 4 . The parallel
transport of Lorentz tensors between tangent spaces on these points, i.e. Tx (M 4 ) 7−→ Tx+dx (M 4 ), is realized through
the spin-connection ω abµ (x), sometimes called Lorentz connection38 . However, we are interested in the differential
forms formalism. Thus, the parallel transport between two near cotangent spaces Tx∗ (M 4 ) 7−→ Tx+dx∗
(M 4 ) is realized
by the Lorentz algebra-valued spin-connection one-form

ω = ω abµ dxµ Ta b . (120)

Hence, due to Ta b = −T ba , we have ω ab = −ωb a . The corresponding exterior covariant derivative is defined in the
usual way,

D =d+ω , (121)

38 The word “spin” rather than “Lorentz” was introduced because the spin-connection is responsible for the coupling with fermions.
29

acting on tangent space indexes. Moreover, the Minkowski local metric ηab and the Levi-Civita tensor abcd are
covariant constant quantities (Exercise 2.10)39 , i.e. Dηab = dηab = 0 and Dabcd = dabcd = 0.
Under Lorentz local transformations, ω behaves as a typical gauge connection,

ω 7−→ L−1 (ω + d)L . (122)

Expression (122) makes the covariant derivative (121) an actual covariant quantity (Exercise 2.11). For infinitesimal
transformations we have (Exercise 2.11)

ω 7−→ ω + Dξ . (123)

Or, in components

ω ab 7−→ ω ab + Dξ ab . (124)

where

Dξ ab = dξ ab + ω ac ξ cb − ω c b ξ ac . (125)

FT
Transformation (122) is a typical gauge transformation. Hence, the spin-connection can be interpreted as the actual
gravity gauge field. The affine connection, having no free frame indexes, is a gauge invariant quantity and is computed
by

Γα α a a b

µν = ea ∂µ eν + ωµb eν . (126)

Expressions (126) and (114) define the change of variables from the first order variables to Palatini variables. In
particular, expression (126) is called consistency condition and is obtained by demanding the vierbein to be fully
covariant constant, i.e.
RA
∇(full)
µ eaν = ∂µ eaν + ω ab ebν − Γα a
µν eα = 0 . (127)

We reenforce the fact that, in the Palatini and first order formalisms, the affine connection is not symmetric
anymore. This should be evident from (126). In fact, the anti-symetric contribution is the definition of the torsion
α
tensor 2Tµν = Γα α
µν − Γνµ .

3. Curvature, torsion and hierarchy identities

The curvature two-form is easely defined by the double action of the covariant derivative (Exercise 2.12)
D

D2 = R = dω + ωω , (128)

where, in components, the curvature two-form reads

Rab = dω ab + ω ac ω cb . (129)

The curvature tensor and the curvature 2-form are related by (Exercise 2.13)
1 a b αβ
Rab = e e R µν dxµ dxν , (130)
2 α β
where Rαβµν generalizes the Riemann tensor (110) for the case of non-vanishing torsion. It is important to emphasize
that (Exercise 2.14) R 6= Dω, which should be clear from (125).
The torsion two-form is defined as the covariant exterior derivative of the vierbein,

T a = Dea = dea + ω ab eb , (131)

39 Hint: Use (125) below.


30

and it is is related to torsion tensor as (Exercise 2.15)


1 a α µ ν
Ta ≡ e T dx dx . (132)
2 α µν
In the case of vanishing torsion, (131) can be solved for ω, namely
ω ab eb = −dea . (133)
By combining (133) and (126) it is possible to show that Γα µν will be Christoffel symbols, as expected (Exercise 2.16).
By the vanishing torsion condition, we see that is possible to describe GR in the first order formalism, in despite of
the fact that gµν is not considered the fundamental field in favor of the vierbein.
The geometrical interpretation of Rab and T a are quite simple. The curvature carries information about the paral-
lelogram rule of vector addition. Hence, by taking the two consecutive parallel transports along different directions,
namely nµ and then tµ . Then take the same vector and parallel transport it again, but inverting the direction order,
namely tµ and then nµ . The resultant vectors will end up at the same point, but will not coincide because the manifold
is curved, i.e. Rab 6= 0. On the other hand, if T a 6= 0, the two resulting vectors may coincide under parallelism, but
will not lie at the same point because the manifold is twisted.
The curvature and torsion obey an hierarchy chain of equations given by the Cartan descent structure equations

FT
(Exercise 2.17):

T a = Dea , (134a)
a
DT = Rab eb , (134b)
DRab = 0 , (134c)

where the last identity one is the Bianchi identities for the curvature. We emphasize here that the set of equations
(134a-134c) are natural consequences of the geometry we are constructing.
RA
4. Einstein-Hilbert action

The change of variables (114) and (126), when applied to the Einstein-Hilbert action (111), leads to (Exercise 2.18)
Z  
1
SEH = d4 x e ab µ ν
ea eb − 2Λ2 + T µa eaµ ,

Rµν (135)
16πG

where e = det(eaµ ) = g and T µa is the projected energy-momentum tensor. The variation of (135) with respect to
a a
eµ and ω bµ leads to (Exercise 2.19),

ab ν 1
eb − eaµ Rνα
bc ν α
eb ec − 2Λ2 = 8πGT aµ ,

Rµν
2
D

a
Tµν = 8πGS abν ebµ , (136)
respectively. In (136), we have defined the spin density as
δT αc c
S abµ = e . (137)
δω abµ α
In the form notation, the Einstein-Hilbert action (135) turns out be an integrated local four-form, polynomial in
the fields and their derivatives, given by (Exercise 2.20)

Λ2 a b c d
Z    
1 ab c d a ab
SEH = abcd R e e + e e e e + Θa e + Sab ω , (138)
32πG 6

where the energy-momentum tensor was split according to Ta ea = Θa ea + Sab ω ab . Clearly, Θa does not depend on
the spin-connection and Sab does not depend on the vierbein. The corresponding field equations (Exercise 2.21) read
now
Λ2 b c d
 
abcd Rbc ed + e e e = 16πGΘa .
3
abcd T c ed = 16πGSab , (139)
31

Needless to say, the first equations in (136) and (139) are equivalent and coincide with the Einstein equations (104)
for the vanishing torsion case. Moreover, since the second equations in (136) and (139) are algebraic, the vanishing
torsion case is always achieved for vanishing spin densities.

C. Gravity as a gauge theory

The similarity between Yang-Mills connection and spin connection should be clear from their gauge transformations,
for instance (33) and (122). Both of them are connections with respect to a gauge group. Moreover, recalling that
the Yang-Mills curvature, in differential form notation, is F = dA + κAA, another similarity appears with respect to
the curvature (128). However, a few important differences are in order:
• The main difference is related to the geometric structures of both theories: The gauge group of gravity has an
intimate relation with spacetime, i.e., the gauge group is identified with the local isometries of the spacetime
tangent space. In Yang-Mills theories, on the other hand, the gauge group describes an internal space with no
relation with spacetime geometry whatsoever. The vierbein is the actual object which connects the internal
group space with spacetime. Hence, the vierbein is the field responsible to address the SEP to the theory. The

FT
consequence is that the corresponding gauge field describes the parallelism properties between near cotangent
spaces.
• Torsion seems to have no analogue with respect to Yang-Mills theories. In fact, a vierbein field is not present
at Yang-Mills theories. Nevertheless, under Lorentz gauge transformations, the vierbein has an explicit matter
character, in despite of its geometrical appeal. Thus, in order to have a parallel, a kind of matter field should
be introduced in Yang-Mills theories.
• The Einstein-Hilbert action (138) has no relation with the Yang-Mills action, which, in differential form notation
reads
Z
1
RA
SY M = FA ∗ FA . (140)
2

• Another point is that the Yang-Mills theory, at least at perturbative level, is massless. The Einstein-Hilbert
action, however, has at least two mass parameters right from the beginning, namely, Newton and cosmological
constants. This feature is not only related to the fact that we have very different actions, but also because the
vierbein is dimensionless with respect to canonical dimensions, i.e. [e] = 0.
• It is also important to notice that the Hodge dual ∗ is not required to construct the EH action. In fact, the
absence of the Hodge operator opens the possibility to interpret gµν as an induced quantity. See next section.
• Finally, by thinking about the Einstein-Hilbert action (138) as a genuine field theory for ea and ω ab , it lacks
quadratic terms in the action, i.e. there are no kinematical therms. Hence, it seems that a field theoretical
D

perturbative approach is out of question, at least in the usual way.

V. GENERALIZED THEORIES

In the first order formalism of gravity there is a limited number of Lagrangians which have physical consistency.
The basic bricks to build a theory of gravity in the first order formulation are ea , ω ab , Rab , T a and the gauge
invariant tensors ηab and abcd . Under the restriction to first order derivatives, there are two main approaches to
obtain generalized actions, namely, the Lovelock action [103] and the Mardones-Zanelli action [93].

A. Lovelock theory

The theory of gravity formulated by D. Lovelock in 1971 consists in to describe the gravitational interaction as an
extension of the Einstein’s GR [103] by adding to Einstein-Hilbert action (138) all possible extra terms but still with
the requirements of locality, vanishing torsion and derivatives order not higher than 1. We notice that, vanishing
32

torsion and first order derivatives imply on second order derivatives for the vierbein, see (133). Lovelock was able to
find a generic action in D dimensions given by
D
2
Z X
SLovelock = ak LkD , (141)
k=0

where ak are generic constants and

LkD = a1 a2 ···aN Ra1 a2 · · · Ra2k−1 a2k ea2k+1 · · · eaD . (142)

For any spacetime dimension, the field equations generated by the action (141) will be second order differential
equations for the metric.
Let us consider, for instance, D = 2. In this case we have (Exercise 2.22)
Z
1
ab a0 ea eb + a1 Rab ,

SL2 = (143)
32πG

FT
which is the usual EH action with cosmological constant in two dimensions. This action is known to be topological,
i.e. it provides no field equations at all (Exercise 2.23).
For D = 4 we obtain (Exercise 2.24)
Z
1
abcd a0 ea eb ec ed + a1 Rab ec ed + a2 Rab Rcd ,

SL4 = (144)
32πG
The first term is recognized as the cosmological constant term while the second term is the well known EH term. The
novelty in (144) is the last term. However, in four dimensions, this term is actually topological and, hence, provides
no contribution to the field equations. This term is known in the literature as the Gauss-Bonnet term.
RA
B. Mardones-Zanelli theory

The Mardones-Zanelli formulation [93, 94] is a generalization of the Lovelock action with the relaxation of the
torsionless condition. A generic expression for the D-dimensional case is not at our disposal. The main new features
at this construction are the inclusion of torsional series and Chern-Simons terms. Let us write a few examples, starting
with the two dimensional case. In two dimensions, no extra terms are added to (143) because all two-form terms that
can be construct are already there.
For three dimensions, on the other hand, we have
Z   
1 a b c ab c
 a a b 2 a b c
SM Z3 = abc a0 e e e + a1 R e + b1 T ea + b2 ω b dω a + ω b ω c ω a . (145)
D

32πG 3

Again, the first term is the cosmological constant term and the second is the EH term. The third term is the Chern-
Simons three-form for the vierbein and the fourth term is the usual topological Chern-Simons three-form for the
connection. It is known that, for certain situations, the last term is equivalent to the first two terms [104].
The four-dimensional case is obtained as
Z
1
abcd a0 ea eb ec ed + a1 Rab ec ed + a2 Rab Rcd + b1 Rab Rab + b2 Rab ea eb + b3 T a Ta .
  
SM Z4 = (146)
32πG
The first three terms coincide with the Lovelock action. The fourth term is the Pontryagin topological term and is
related to the connection CS term through
 
2
Rab Rab = d ω ab dω ba + ω ab ω bc ω ca . (147)
3

The last two terms are, in fact, equivalent up to surface terms. Moreover, for b2 = −b3 , the last two terms become
the Nieh-Yang topological term which is related to the vierbein CS term [105],

Rab ea eb − T a Ta = d(T a ea ) . (148)


33

VI. QUANTIZATION ATTEMPTS

It is widely known that GR has deep problems in what concerns its quantum consistency. In fact, the first attempts
to quantize GR were found to be inconsistent [130–132]. The idea is to perturb the metric around a fixed background40 ,
for instance, the Minkowski background,

gµν = ηµν + khµν , (149)

where k is a constant with negative mass dimension [k] = −1 related to Newton‘s constant, and quantize the gravita-
tional waves modes hµν . The linearized version of the EH action (111) is known as the Pauli-Fierz action and enjoys
an Abelian gauge symmetry for the field hµν . The quantum field hµν , being symmetric in its indexes, describes spin-2
excitations associated to the graviton field. However, the standard quantization methods for gauge theories fail due to
the fact that the theory is not renormalizable, i.e. extra couplings are required. The extra counterterm indicates that
quadratic terms on the curvature should be included in the original theory. If this is correctly done, then the linearized
theory is actually renormalizable. However, the extra terms provide non-physical modes associated to higher order
derivatives [133]. Consequently, although renormalizable, the theory with higher order derivatives is not unitary.
Many other alternatives to quantize GR have been proposed. Perhaps, the closer approach to the background

FT
linearization method above described is the asymptotic safety method [22, 23]. In this method, the renormalization
group method is employed in the looking for fixed points, in despite of the non-renormalizability of EH action. The
claim is that EH action is non-perturbative, hence, perturbative expansions and counterterms absorption may not
make sense. In this sense, the full non-linearized EH action is considered in a path integral approach. Of course, to
find these fixed points is a highly non-trivial task and the final solution still lacks.
Another popular technique is the loop quantum gravity theory [20, 21] where a different set of gravity variables
is quantized as fundamental quantities. These variables are related to the spacetime holonomies and, hence, auto-
matically predicts spacetime granulation. the method is background independent and it seems to possess evidence
of a nice classical limit [136]. However, huge obstacles remain to be overcame such as the solution of the so called
Hamiltonian constraint and the understanding of how time is encoded in this theory.
RA
More recently, a power counting renormalizable proposal have emerged, the Horava-Lifshitz gravity [24–26]. In this
approach, spacetime diffeomorphisms are explicitly broken and power counting renormalizable actions are constructed
with space and time independent terms. The diffeomorphisms are assumed to be a macroscopic phenomenon, i.e. all
parameters of the theory should converge to macroscopic values for which diffeomorphism is gained. Of course, the
formal renormalizability proof lacks and a consistent classical limit is not known.
A different approach to treat quantum gravity is the large class of theories known as emergent gravities. In these
theories, gravity as geometrodynamical theory is an emergent phenomenon produced by all kinds of quantum effects
which have nothing to do with spacetime at some energy scale. In other words, gravity as geometry is a classical
limit of a more fundamental theory which is not related to spacetime dynamics. Hence, one could say that quantum
gravity and geometrodynamics would be different sides of the same coin. some examples we will not discuss can be
found in [27–30] and references therein.
Perhaps the most popular method to describe gravity in quantum level is superstrig theory [16–19] which concerns
D

the quantization of string systems instead of particles. In this approach, all spin excitations are contained in the string
vibration modes, including the spin-2 graviton modes. Hence, gravity as geometrodynamics emerges as an effective
effect. Moreover, superstring theory is known to have the property of unifying all four interactions. Nevertheless,
it predicts extra particles still not observed by experiments such as the superpartner particles. It also predicts the
spacetime dimension to be ten, i.e. six more than the four known dimensions. Just like the extra particles, the extra
dimensions have not been observed yet.
Still in effective gravity theories, there is a whole bunch of gauge theories which could lead to geometrodynamics, see
for instance [31–42]. Generally speaking, the idea is to define a metric independent action enjoying gauge symmetry.
In most approaches, a topological action is defined and a spontaneous symmetry breaking is enforced, freeing the
degrees of freedom which will be interpreted as gravity. An example in which the starting point is not a topological
action and there is no need of Higgs fields is explored with some detail in the next part.

40 The same idea can be employed in the first order variables.


34

PART III: EFFECTIVE GRAVITY FROM YANG-MILLS THEORIES

In this part we focus on a specific mechanism that generates a generalized geometrodynamical gravity theory from
a Yang-Mills action based on the group SO(5). A deep analogy with QCD is discussed. As so, the Yang-Mills
action is taken as a proposal for a quantum gravity theory while the geometrodynamical classical limit would be a
generalization of GR. These and similar ideas can be extensively found in [31–42] and references therein.
In most of this part we will remain at classical level, at least for most of the technical discussions. Nevertheless,
the idea is that everything we do with SY M should be extended to its renormalized quantum version Γ.

VII. GENERAL IDEAS

As discussed in Sec. VI, geometrodynamical theories such as gravity have many problems in what concerns their
quantization. In fact, the principles of QFT and GR seem to carry fundamental incompatibilities. The main obstacle
in quantizing a system is to find, if any, the correct quantization method because the classical theory is always a limit
of another quantum theory. In this sense, the quantization techniques can not be taken as fundamental approaches
to obtain the quantum version of a classical system: The most fundamental theory is the quantum one while the

FT
classical theory is a particular macroscopic case, i.e. an effective version of the quantum theory. The Faddeev-Popov
quantization and Gribov ambiguities are striking examples of that idea.
Remarkably, in some cases, the quantum and classical theories seem to have no relation whatsoever. This is
the case of the SM, which is fully described by Yang-Mills theories coupled to spinorial (quarks and leptons)
and scalar (Higgs) fields (needless to say, Yang-Mills theories displays quantum consistency, renormalizability and
unitarity). In fact, at very high energies, the SM is a gauge theory based on the group SU (3) × SU (2) × U (1)
where the corresponding gauge fields are massless and mediate the interactions between all matter massless
fields. As the energy decreases, the three sectors split, with different behavior each. The QCD sector is confined
and develop several mass parameters (Gribov mass, condensates...) and suffers hadronization while the elec-
troweak sector suffers spontaneous symmetry breaking to the subsector U (1): roughly speaking, the vector bosons
RA
become massive while the photon remains massless. Hence, at low energies, three completely different theories emerge.

The question we can formulate is: Can gravity also be an exotic limit of a typical gauge theory?
Or, equivalently: Can quantum gravity be a standard gauge theory in four dimensions?

The answer is not trivial... at all: In all three fundamental interactions of the SM, the path between the
perfect Yang-Mills theories at high energies and the cumbersome low energy regime is quite difficult, even obscure.
In the case of QCD, the hadronization phenomenon, intimately linked to confinement, lacks a theoretical description.
In other words, how to start from the Yang-Mills action and end up at an action in terms of hadrons and glueballs?
The essence of this question almost perfectly fits to our gravity problem. In QCD we know the quantum theory and
the final low energy states. In gravity, we know the effective low energy theory, i.e. GR. However, we do not know
the starting fundamental theory, nor the mechanism leading it to geometrodynamics.
D

Therefore, as a starting point, because Yang-Mills gauge theories work so nicely in flat spacetime for three of the
four fundamental quantum interactions, we postulate that:

Quantum gravity is a Yang-Mills gauge theory in four-dimensional Euclidean spacetime R4 .

This choice, instead of a Minkowski spacetime, has three main reasons:

• First, because QFT is only solvable in Euclidean spaces. Reliable computations require a Wick rotation from
the Minkowkian to the Euclidean space [6]. However, a Wick rotation is known to be valid only at perturbative
level. At non-perturbative scales, there are no indications that a Wick rotation can be performed. Thus, since
we are interested in non-perturbative effects, it is necessary to start with a Euclidean spacetime.
• The second reason, and perhaps the most important, is that the Euclidean space is the simplest geometry. Since
we are constructing a theory that will determine spacetime geometry, starting with the simplest geometry seems
to be the right choice.
• Third, in four dimensions, the coupling parameter of a Yang-Mills theory is dimensionless. As a consequence,
35

pure Yang-Mills action is massless41 . This fact is important because it prevents the gauge field, which has
dimension 1, to be directly associated with the vierbein, which has vanishing dimension. This means that, at
quantum level, the gauge theory and the spacetime are completely unlinked.

Hence, from the third motivation above, to identify the vierbein at low energies, it seems that mass parameters
must appear at some level. Moreover, since Newton’s constant carries dimension −2, it becomes evident that at least
one mass parameter must appear in order to obtain gravity at low energies. Therefore, since Yang-Mills theories
naturally generate mass parameters as the energy decreases, our choice sounds promising. In fact, we pick the Gribov
parameter as the mass that will be responsible for this generalization.
It remains to decide which group (or class of groups) is the most suitable group to describe our mechanism. The
gauge field is an algebra-valued 1-form and thus, carries dim G components with respect to the algebra of the gauge
group. The gauge group G has to fulfill a few requirements [39, 107]:
• The group must have dim G ≥ dim ISO(4), in such a way that the degrees of freedom of gravity are covered.
• The group must decompose at least42 as G = H + Q where Q = G/H must be a symmetric space and H,
obviously, a stability group. The stable group must share a morphism with Lorentz-type groups and Q must
define a vector representation of H. Thus, H can be identified with local Lorentz transformations and Q with

FT
a sector which expands the vierbein. As a consequence, the gauge components can be identified with the
spin-connection and the vierbein.
• It can be very convenient to have a symmetry breaking G → H, in such a way that the field at the Q sector
acquires a matter-type transformation with respect to H. That is why Q must be a symmetric space. It is
important to mention that, depending on the gauge group, a symmetry breaking might not be needed. This
occurs if the gauge transformations could be partially identified with local Lorentz transformations and partially
identified with diffeomorphisms. A renowned three dimensional example can be found in [104, 108].
Perhaps the simplest group satisfying these conditions is the (Exercise 3.1) N = 5 special orthogonal group. Then,
the choice is that G = SO(5) is the symmetry group of our Yang-Mills gauge theory describing quantum gravity. In
RA
summary, gravity will be described at quantum and classical level as follows:

The theory has two sectors, the UV, which is a massless gauge theory in Euclidean four-dimensional space,
and the IR sector, which presents soft BRST symmetry breaking and dynamically generated mass parameters. The
UV sector is a standard, non-Abelian, asymptotic free, gauge theory of spin-1 excitations. Although the degrees
of freedom coincide in number with a first order gravity, this identification is forbidden unless a mass parameter
arises, so that a vierbein can be defined. Once the energy starts to decrease, the soft BRST symmetry breaking
takes place, the Gribov parameter and other possible masses appear. At this stage, propagators of the fundamental
fields develop complex poles, a fact that is interpreted as an evidence that these excitations are ruled out from the
physical spectrum of the theory (in QCD, this is recognized as confinement). Physical observables must be defined
at this point. In QCD, the observables are hadrons and glueballs. In gravity, the low energy observables must be
D

geometry. Thus, one possibility is to identify the H gauge field with the spin-connection and the Q field with the
vierbein. The consequence is a first order gravity where geometry is determined by the usual relations (114) and (126).

If all above requirements are fulfilled, we gain a deep analogy between gravity and quantum chromodynamics.
At high energies, both theories are well defined quantum gauge theories in a four-dimensional flat space. Both
theories present soft BRST breaking and have to be redefined in order to establish the physical content at the
infrared regime. In the case of QCD, the physical content are confined states identified with hadrons and glueballs.
In gravity, the physical content are identified with geometric properties of spacetime.
We now turn to the specifics of these ideas.

41 In Yang-Mills theories, the four-dimensional case is the only case where the coupling parameter is dimensionless.
42 Larger groups are also allowed. The extra decompositions will generate a matter sector in the resulting gravity theory, [40].
36

VIII. SO(5) YANG-MILLS THEORY

As discussed, one possible realization of the above ideas occurs in a gauge theory based on the group SO(5) in a
four-dimensional Euclidean spacetime, R4 . The algebra of the group is given by
 AB CD  1  AC BD
+ δ BD J AC − δ AD J BC + δ BC J AD ,
 
J ,J =− δ J (150)
2
where J AB are the 10 anti-hermitian generators of the gauge group, antisymmetric in their indexes. Capital Latin
indexes are chosen to run as43 {5, 0, 1, 2, 3}. The SO(5) group defines a five-dimensional flat space, R5S , with invariant
Killing metric given by δ AB ≡ diag(1, 1, 1, 1). The spaces R4 and R5S have no dynamical relation whatsoever.
The SO(5) Yang-Mills action is renormalizable and unitary. Thus, Gribov ambiguities and BRST soft breaking
takes place. Moreover, a few extra mass parameters may emerge. The presence of these masses will be forwardly
used. However, we will fix our attention only at the Yang-Mills action and the gauge field one-form, here denoted by
Y.

A. Projection and rescaling

FT
The five-dimensional orthogonal group may be decomposed as a direct product, SO(5) ≡ SO(4) × S(4) where
S(4) ≡ SO(5)/SO(4) is a symmetric coset space with four degrees of freedom. This decomposition is carried out by
projecting the group space in the fifth coordinate A = 5. Defining then J 5a = J a , where small Latin indexes vary as
{0, 1, 2, 3}, the algebra (150) decomposes as (Exercise 3.2)
 ab cd  1  ac bd
δ J + δ bd J ac − δ ad J bc + δ bc J ad ,
 
J ,J = −
2
 a b 1
J ,J = − J ab ,
2
RA
 ab c  1 ac b
δ J − δ bc J a ,

J ,J = (151)
2
where δ ab ≡ diag(1, 1, 1, 1), J ab ∈ algSO(4), and J a ∈ algS(4). The gauge connection follows the same decomposition,
Y = Y AB JA B = Aab Ja b + θa Ja where, obviously, A ∈ algSO(4), and θ ∈ algS(4). The gauge transformations in
Eq. (34) are decomposed, at infinitesimal level, as (Exercise 3.3)
κ a
Aab 7−→ Aab + Dαab − (θ ζb − θb ζ a ) ,
4
θa 7 → θa + Dζ a + καab θb .
− (152)
where the full gauge parameter splits as ξ = αab Ja b a
+ ζ Ja and D = d + κA is the covariant derivative with respect
to the sector SO(4) of the algebra. The field strength also decomposes (Exercise 3.4),
D

 κ 
F = Ωab − θa θb Ja b + K a Ja , (153)
4
where
Ωab = dAab + κAac Acb ,
K a = Dθa = dθa + κAab θb . (154)
The starting action is, as postulated, the Yang-Mills action (39). In form notation, it reads (Exercise 3.5)
Z
1
SY M = F AB ∗ FA B , (155)
2
which, from (153), decomposes as (Exercise 3.6)
Z  
1 1 1 1
S= Ωab ∗Ωab + K a ∗Ka − Ωab ∗(θa θb ) + θa θb ∗(θa θb ) . (156)
2 2 2 16

43 In comparison with the notation used in PART I, the correspondence is given by A (PART I) 7−→ AB = −BA (PART III).
37

Once the energy starts to decrease the set of mass parameters, together with BRST soft breaking, dynamically
arises. At this point, it is convenient to perform the following redefinitions
7 → κ−1 A ,
A −
θ −7 → κ−1 mθ , (157)
where m is a mass scale which depends on the mass parameters of the theory and still has to be determined.
The transformations (157) are not accidental. Both sectors are rescaled with κ−1 in order to factor out the coupling
parameter outside the action, a standard procedure in Yang-Mills theories [6]. On the other hand, the mass parameter
affects only the θ-sector, transforming it in a field with dimensionless components. It turns out that this is the unique
possibility if one wishes to identify θ with a vierbein field. If also A is rescaled with a mass factor, then it would never
be possible to identify it with the spin-connection. The action (156) is then rescaled to (Exercise 3.7)
m2 a m2 a m4 a
Z  
1 a b b b
S= 2 Ω b ∗Ωa + K ∗K a − Ω b ∗(θa θ ) + θ θb ∗(θa θ ) , (158)
2κ 2 2 16
a a
where Ω b = dAab + Aac Acb , K = Dθa , and the covariant derivative is now given by D = d + A. Moreover, a
reparameterization of the SO(5) generators is required due to the existence of a mass scale, i.e., a stereographic
projection is now allowed if one identifies the mass parameter with the radius of the gauge manifold R5S , see [39].

FT
In essence, the whole θ field must be invariant under the rescaling (157), i.e., under (157), θa Ja 7−→ θa Ja . The
consequence for the algebra (151) is (Exercise 3.8)
 ab cd  1  ac bd
δ J + δ bd J ac − δ ac J bc + δ bc J ad ,
 
J ,J = −
2
 a b m2
J ,J = − 2 J ab ,

 ab c  1 ac b
δ J − δ bc J a .

J ,J = (159)
2
So far, we have done nothing. Action (158) is exactly the original action (155) with a specific decomposition of the
RA
fields and a very simple rescaling of the fields.

B. Contraction and symmetry breaking

In order to obtain gravity, the main step is to consider that the rate m2 /κ2 tends to vanish at a specific low energy
scale44 . At this point, this is only an assumption, however, due to asymptotic freedom, it is not an improbable
hypothesis45 . Taking this limit at the rescaled algebra (159), an Inönü-Wigner contraction46 [109] takes place,
enforcing the SO(5) group to be contracted down to the Euclidean Poincaré group (Exercise 3.10). The gauge
symmetry is then dynamically deformed to the Euclidean Poincaré group, SO(5) −→ ISO(4), for some values κ
in the strong coupling regime. The details of this deformation can be understood as a stereographic projection
D

(Exercise 3.11), see the appendix in [42]. An Inönü-Wigner contraction is a deformation of the group to another
group which is not a subgroup of the former. Hence, since the Poincaré group is not a symmetry of the action (158),
a dynamical symmetry breaking takes place. Nevertheless, the Lorentz-type group SO(4) is not only a common
subgroup, ISO(4) ⊃ SO(4) ⊂ SO(4), but is also a stability subgroup for both groups. Thus, the theory suffers
a symmetry breaking SO(5) −→ SO(4). Under the SO(4) gauge symmetry, the transformations (152) reduce to
(Exercise 3.12)
Aab −
7 → Aab + Dαab ,
θa −7 → θa − αab θb , (160)
where (157) was assumed. Thus, after the contraction, the field A becomes the only gauge field with respect to the
Lorentz group while θ has migrated to the matter sector (it suffers only group rotations under infinitesimal Lorentz
transformations, i.e. it is a vector representation of the group). Now, the theory is ready to be identified with gravity.

44 We remark that, due to the hierarchy of fundamental interactions, for gravity, a low energy scale would be much higher when compared
to all other fundamental interactions.
45 In Sec. X C we will show that this is actually what happens when we take m to be the Gribov mass parameter.
46 Perhaps the most known example of an Inönu-Wigner contraction in Physics is the deformation of the Lorentz group to the Galileo
group for v/c 7−→ 0 (Exercise 3.9).
38

IX. INDUCED GRAVITY

The broken theory described in Sect. VIII B can generate an effective geometry if [39, 40, 107]: (i) every configuration
(A, θ) defines an effective geometry (ω, e); (ii) there exists a mapping from each point x ∈ R4 to a point X ∈ M4
of the deformed space. In order to preserve the algebraic structure already defined in R4 , it is demanded that this
mapping is an isomorphism; (iii) The local gauge group SO(4) defines, at each point of the mapping, the isometries
of the tangent space TX (M). Thus, θ and A can be identified with the vierbein e and spin connection ω, respectively,
through

ωµab (X)dX µ = δaa δbb Aab µ


µ (x)dx ,
eaµ (X)dX µ = δaa θµa (x)dxµ . (161)

In expressions (161), Latin indexes {a, b . . .} belong to the tangent space TX (M). Moreover, it is always possible
to impose that the space of all p-forms in R4 is mapped into the space of p-forms in M4 , namely, Πp 7−→ Π e p , and
the same for the Hodge spaces, ∗Πp 7−→ ?Π e p , where ? is the Hodge dual in M4 . This mapping47 , together with the
identifications in Eq. (161), provides (Exercise 3.12)

FT
Λ2
Z  
1 1 a 1
S= 2
R b ? Rab + T a ? Ta − abcd Rab ec ed + abcd ea eb ec ed , (164)
16πG 2Λ 2 4

where Rab = dω ab +ω ac ω cb and T a = dea +ω ab eb are the curvature and torsion, respectively. Newton and cosmological
constants are determined by

m2 = κ2 /4πG = 4Λ2 . (165)

Action (164) is a gravity action in the first order formalism presenting (in order of appearance) a quadratic Yang-
RA
Mills-type curvature term, a quadratic torsion term, the Einstein-Hilbert term, and the cosmological constant term.
In summary, an Yang-Mills theory based on the SO(5) group can be identified with gravity if a mass parameter
emerges at some energy scale and if the 2-step deformation SO(5) 7−→ ISO(4) 7−→ SO(4) takes place. In our
hypothesis, the deformation is driven by a dynamical behavior of the ratio m2 /κ2 7−→ 0. In a way, the degrees of
freedom of the gauge fields are absorbed by spacetime which acquires a deformed character. Hence, all other matter
fields would feel this deformation in the way, independently of their character. This phenomenon can be visualized
as the rising of the equivalence principle.
Now, we start to explore the physical contents of this model at quantum and classical levels.

X. QUANTUM ASPECTS
D

A. Weinberg-Witten theorems, emergent gravity and spin-1 gravitons

In [110], Weinberg and Witten established two very powerful theorems in QFT. Essentially, they forbid: (i) massless
charged states with helicity j > 1/2 which have a conserved Lorentz-covariant current and (ii) massless states with
helicity j > 1 which have conserved Lorentz-covariant energy-momentum tensor.

47 The formal aspects of the mapping between a gauge theory in Euclidean spacetime and a gravity theory can be discussed in terms of fiber
bundle theory. The details can be found in [39, 40, 107]. The result was discussed for general manifolds with d arbitrary dimensions.
Let the original manifold have metric tensor gµν (x) while the effective metric tensor is defined as g
eµν (X). Thus, the matrix that defines
the map x 7−→ X for all points in the effective manifold is given by
 1/2d

Lν µ = g̃ να gαµ . (162)
g
Its inverse is given by
 1/2d
ν g
L−1 µ
= g να g̃αµ . (163)

In (162) and (163), g = det gµν and g
e = det g
eµν are taken as non-vanishing quantities. Thus, ambiguities are absent in this mapping.
39

At first sight, the first theorem would forbid gauge theories to exist because vector bosons and gluons are charged
massless states which carry helicity j = 1. However, the conserved gauge current associated with these states are
not Lorentz covariant [110]. Similarly, the second theorem forbids spin-2 states with conserved Lorentz-covariant
energy-momentum tensor to be defined. In traditional GR the metric tensor gµν is not a Lorentz covariant tensor
since it is constructed over a Riemannian manifold. The linearized version of GR, on the other hand, is also a gauge
theory and the same principles of the first case apply [110]. Moreover, the energy-momentum tensor of the graviton
field identically vanishes.
On the other hand, the Weinberg-Witten theorem (ii) applies to emergent gravity theories for which massless spin-2
states are generated as effective or composite states that could be associated with graviton excitations. Nevertheless,
one may argue that the Weinberg-Witten theorem (ii) applies to the present mechanism. However, this is not the
case here. First, the theory has a few mass parameters and the theorem holds only for massless states. Second, and
more important, there are no spin-2 states in this model. The fields are identified with geometry and not with spin-2
composite fields. Gravity emerges as geometrodynamics and not as a field theory for spin-2 particles in flat space.
Nevertheless, renormalizability establishes that the quantum action has the same form of its classical version. The
difference between the classical and quantum actions lies on the fields and parameters, which are their respective
renormalized versions. Hence, although each configuration (A, θ) can define a geometry (ω, e), the mapping should
not be applied to each configuration in the path integral, but at the quantum action itself. In this way, the resulting

FT
geometrodynamical theory is obtained from the full dynamical content of the original gauge theory. The conclusion
is that there is no violation of the Weinberg-Witten theorem.
Another question may raise at this point: Where are the spin-2 excitations of the graviton? And it is an important
question. The main point here was already discussed: The present approach is not a quantization of gravity geometric
variables. Inhere, gravity is an effective manifestation of another theory, i.e. geometrodynamics is a classical limit
of a quantum gauge theory of spin-1 excitations. On the other hand, if our SO(5) gauge theory is in fact quantum
gravity and the graviton is the mediator of gravity at quantum level, then, in the present theory, the graviton has
spin-1. Hence, this definition differs from the usual graviton definition which is associated with massless spin-2 states.
Nevertheless, after the emergence of gravity as geometrodynamics, linearization is allowed for weak gravitation and
the spin-2 states might also be considered. In that case, these states are classical states associated with propagating
RA
fields and not quantum states.
Another way to think of the spin counting difference lies on the map R4 7−→ M4 where the local gauge group
is identified with the isometries of the tangent spaces of M4 . In this process, the group indexes are interpreted as
spacetime tangent indexes (an association prevent at high energies due to the massless character of Yang-Mills action).
Thus, after the map, each field gains an extra index and thus, extra spin degrees of freedom.

B. Cosmological constant problem

A remarkable feature of the present theory is that Newton and cosmological constants are related through Λ2 =
κ2 /16πG. From asymptotic freedom, κ2 is a big quantity at low energies. And, by assumption, G is small. Thus,
Λ should be very big. In fact, if this is true, we can make two important remarks: (i) The first term in (164) can
D

be safely neglected. Moreover, due to the absence of matter fields, torsion can be taken as very small. The resulting
theory is then, the usual Einstein-Hilbert theory with (a very big) cosmological constant (see Sec. XI A); (ii) although
astrophysical predictions [111, 112] determine that Λ2obs is very small, quantum field theory predicts [113–115] a very
large Λ2qf t . Thus, the contribution of a pure gravitational cosmological constant, which is big in our case, can drive the
cosmological puzzle to a final consistent answer. In fact, following [39, 116], the renormalized cosmological constant
of our model could be determined through Λ2obs = Λ2ren + Λ2qf t ≡ Λ e 2 . This means that, by summing all vacuum
contributions to the last term in (164), it should be replaced by Λ e 2 . We will in fact assume this at Secs. XI C and
XI B.

C. Quantum predictions

It is possible to perform actual explicit computations within a semi-perturbative approach, see [40] for the first
1-loop estimates and [117] for 1 and 2-loop improved computations. In these notes, without loss of generality, we
confine ourselves to 1-loop results.
The important thing at this point is to wisely find the correct mass parameters to account for m. It is intuitive
that the Gribov parameter is the most probable candidate because it is a genuine non-perturbative and inevitable
parameter and it is gauge invariant [49, 50].
40

1. Running parameters

In perturbative QFT, the running of the parameters can be determined by the renormalization factors of fields and
parameters. In the case of the Gribov parameter, it is determined by a gap equation obtained from the minimization
of the quantum action with respect to γ 2 , see Sec. III. At 1-loop, the gap equation is48 [14, 46]
3N κ2 d4 p
Z
1
=1. (166)
4 (2π)4 p4 + γ 4
where N = 5 and p is the internal momentum. Following [46], by employing the MS renormalization scheme at the
gap equation (166), it is not difficult to find (Exercise 3.13)
N κ2 5 3 N γ4
  
− ln =1, (167)
16π 2 8 8 µ4
where µ is the energy scale and the trivial solution γ 2 = 0 was excluded. Equation (167) provides (Exercise 3.14)
5
e6 16π 2
 
−4
γ 2 = √ µ2 e 3 N κ2 . (168)

FT
N
On the other hand, recalling that the coupling parameter is determined through the renormalization group equations
[66, 67], we can write

N κ2 1
= µ2
, (169)
16π 2 11
ln 2
3 Λ

where Λ is the renormalization group cutoff. Hence, combining (168) and (169), it is found (Exercise 3.15)
−35/9
RA
5
µ2

e6 2
γ2 = √ Λ 2 . (170)
5 Λ
We can see from (170) and Fig. 4 that, as higher the energy is, as smaller the Gribov parameter is. This is the expected
behavior [14, 46, 48] of γ 2 because at high energies perturbative massless Yang-Mills theory must be recovered. At
the deep infrared region, however, it seems to diverge. This behavior is an evidence that the semi-perturbative
approximation needs improvements. In fact, it is known from lattice simulations for unitary groups that the coupling
parameter is finite at the origin [118, 119]. This kind of improvement could affect the infrared sector of the Gribov
parameter in such a way that it could also be finite.
D

5
e 6 2
FIG. 1: Gribov parameter as function of energy scale. The energy is in units of Λ and the Gribov parameter in units of √ Λ .
5

48 It is important to notice that, because we are dealing with Yang-Mills theories, the following computations can be derived in the same
manner that of [46], only keeping in mind that the fundamental Casimir here is 5 and the group dimension is 10.
41

Now, the ratio γ 2 /κ2 is easily determined from (169) and (170),

5
! −35/9
γ2 µ2 µ2
  
55 e6 2
= √ Λ 2 ln 2 . (171)
κ2 48π 2 5 Λ Λ

The very interesting behavior of γ 2 /κ2 is plotted in Fig. 2 and the inverse ratio is displayed in Fig. 5. As expected,
at high energies, γ 2 /κ2 asymptotically vanishes. Then, in a scale right above Λ, γ 2 /κ2 achieves a maximum. At this
region, BRST soft breaking dominates and the rescaling of the fields (157) is allowed. After that, it fastly drops to zero
at µ = Λ. Is exactly at this point that the theory suffers the Inönü-Wigner deformation which induces the breaking to
the SO(4) theory and the geometric phase starts over. This point is also recognized as the point of phase transition
in non-Abelian gauge theories. Below this point, another theory takes place. In the case of QCD, ΛQCD ≈ 237M eV ,
the Yang-Mills action should be replaced by an action based on hadrons and glueballs excitations. In the case of
gravity, the Yang-Mills action must be substituted by a geometrodynamical action and the phase transition scale is
expected to be around Planck scale.

FT
RA
2 2
FIG. 2: The
 ratio γ /κ as function of energy scale. The energy is in units of Λ and the Gribov parameter in units of
5
55 e 6 2
48π 2

5
Λ .

Below the transition point Λ, where the coupling parameter diverges, the squared coupling parameter acquires
negative values. This indicates that below Λ the perturbative predictions are actually meaningless. However, from
lattice predictions [119, 120], there are strong evidences that the non-perturbative coupling is actually finite at the
origin and presents no divergence at the transition scale. On the other hand, we can argue that the divergence is a
D

strong signal that there is a phase transition at that point. Thus, neither way, the coupling parameter should drop out
in favor of an effective coupling. In the case of QCD, it is not known how to obtain this parameter from the dynamics
of Yang-Mills theories. However, in the present case, we can interpret the ratio γ 2 /κ2 as the effective coupling, which
is related to Newton‘s constant.
The analysis for the cosmological constant is easier because Λ ∼ γ 2 . Thus, it should be very large indeed. As
discussed in Sec. X B, this is actually a very welcome feature because it may compensate the QFT predictions and,
eventually, it can provide a small effective value [116].

2. Numerical estimates

There are two main ways to calculate the parameters of interest. The usual way [6, 68] is: i) to fix the renormalization
group parameter at the transition energy. Its value is a phenomenological quantity and cannot be derived by theory. In
2
our case is Planck energy; ii) chose a energy scale in such a way that N κ2 /16π 2 and ln(µ2 /Λ ) are as small as possible;
iii) compute the physical parameters you want. Another way is: i) Fix a physical parameter to its experimental value;
2
ii) chose a energy scale in such a way that N κ2 /16π 2 and ln(µ2 /Λ ) are as small as possible; iii) Compute the other
parameters as well as the transition scale. We choose to discuss here the second way (both ways are discussed in
detail in [117] and there are no significant difference between them).
42

2 2
FIG. 3: The
 ratio κ /γ as function of energy scale. The energy is in units of Λ and the Gribov parameter in units of

48π 2
5 −2
Λ 2.

FT
55 5
e6

Following the second approach, the scales µ and Λ can be estimated by fixing the current value of Newton’s constant
to G−1 ≈ 1.491 × 1032 T eV 2 . Thus, combining (165) and (171), one easily achieves
 −70/9  2 
31 µ2 2 µ
7.323 × 10 T eV = Λ ln 2 . (172)
Λ Λ
2
We have little freedom to choose µ as long as N κ2 /16π 2 < 1. Let us work at µ2 = 2Λ . Thus, N κ2 /16π 2 ≈ 0.393.
RA
2
This provides ln(µ2 /Λ ) ≈ 0.693. Accepting these as reasonable values, we achieve for the renormalization group scale
2
Λ ≈ 2.122 × 1033 T eV 2 , (173)

which is quite close (for a 1-loop semi-perturbative estimate) to Planck energy Ep2 = 1.491×1032 T eV 2 . Although some
approximations and extrapolations have been considered, we can interpret these values as a good result, indicating
that the geometric phase of gravity appeared right before Planck scale. This means that right above Planck scale,
where quantum mechanics starts to make sense, gravity is already in its geometric phase.
We can also estimate the value of the cosmological constant for the chosen scales. From (165) and (170), it provides

Λ2 ≈ 1.106 × 1032 T eV 2 . (174)


D

This is a huge amount of energy and is three orders of magnitude greater than quantum field predictions [121] of
Λ2qf t ∼ −3.71 × 1028 T eV 2 and would not cancel it to provide the observational data value Λ2obs ∼ 1.686 × 10−68 eV 2 .
However, for 1-loop approximation at zero temperature, it is quite remarkable that a solution accommodating Newton
constant and Planck energy scale could be found.

XI. CLASSICAL ASPECTS

Let us now study some classical aspects of the theory.

A. Classical field equations

The variation of the action (164) with respect to ω and e leads to (Exercise 3.16)

Λ2 b c d
 
3 bc b bc d
R ? (Rbc ea ) + T ? (Tb ea ) + D ? Ta − εabcd R e − e e e =0, (175)
2Λ2 3
43

for the vierbein, and (Exercise 3.17)

3
D ? Rab + 2eb ? Ta − εabcd T c ed = 0 , (176)
Λ2

for the spin-connection. The last term in (175) is recognized as the Einstein term with cosmological constant.
Nevertheless, as discussed in Sect. X B, the cosmological constant term in (175) will sum with all vacuum terms of the
matter fields action. Therefore, by this phenomenological argument, we should replace Λ 7−→ Λ, e only at this term.
Thus,
!
3 bc e2
Λ
R ? (Rbc ea ) + T b ? (Tb ea ) + D ? Ta − εabcd bc d
R e − eb ec ed =0, (177)
2Λ2 3

where Λ e is the observational cosmological constant, which is extremely small. Equations (175) and (176) are the full
field equations, valid at all sectors of the geometric phase. On the other hand, for a sufficient low energy scale where
all other theories achieve their vacuum states, equations (177) and (176) must be employed.

FT
The Einstein limit can be obtained with a very subtle argument: Since Λ2 is a very large quantity, all terms
proportional to Λ−2 are small perturbations as long as the curvatures are not too strong. Hence, at zeroth order
approximation, equations (177) and (176) reduce to
!
e2
Λ
b bc d b c d
T ? (Tb ea ) + D ? Ta − εabcd R e − e ee = 0.
3
2eb ? Ta − εabcd T c ed = 0 . (178)

The second of (178) is algebraic in the torsion and, hence, a solution T = 0 is allowed49 . Thus, (178) reduce to
RA
!
e2
Λ
εabcd Rbc ed − eb ec ed = 0 . (179)
3

which is the well know Einstein equations, i.e. the first of (139) in vacuum.

B. Spherical symmetry static solutions

Let us consider the field equations (176) and (177) for vanishing torsion and very high values of Λ2 ,
!
D

2
3 bc Λ̃
R ? (Rbc ea ) − εabcd Rbc ed − eb ec ed = 0 . (180)
2Λ2 3

We remark here that the first term in (180) will be considered as a very small perturbation. In, (176), for vanishing
torsion, the remaing term is virtually zero and can be safelly neglected since it does not affect any other dynamical
term.
We take the usual ansatz [10–13] for static and spherically symmetric metric, namely

e0 = −eα(r) dt , e1 = eβ(r) dr , e2 = rdθ , e3 = r sin θdφ . (181)

49 It is important to understand that this limit can only be taken outside a matter distribution of for sufficient small spin densities,
otherwise, T 6= 0 in general.
44

Hence, equation (180) decomposes as


"  2  2 #
e−2β ∂r β 1 − e−2β
 −2β
1 − e−2β

e ∂r β
σ 2 + + 2 + + 3λ = 0 ,
r r2 r r2
"  2  2 #
e−2β ∂r α 1 − e−2β
 −2β
1 − e−2β

e ∂r α
σ 2 + 2
− 2 + + 3λ = 0 ,
r r r r2
i2  e−2β ∂ α 2  e−2β ∂ β 2
( )
h
r r
σ e−(α+β) ∂r e−β ∂r eα

+ + +
r r
 e−2β ∂r α e−2β ∂r β
−e−(α+β) ∂r e−β ∂r eα − + + 3λ = 0 ,
r r
(182)

for a = 0 , a = 1 and a = 2, respectively. The constants in (182) are


3

FT
σ = − ,
2Λ2
Λ̃2
λ = − , (183)
3
We notice that the differential equations for a = 2 and a = 3 are identical. The remaining equations can be solve by
perturbation theory around the well-known de Sitter Schwarzchild solution []. A long computation yields, at fourth
order in perturbation theory,
 
−2β a1 2 b1 b3 c
1 c3 c4 
e = 1+ + a2 r − η 2
+ b2 r + 4 − η 2 + c2 r2 + 4 + 7 +
r r r r r r
RA
 
d1 d3 d4 d5  e1 e3 e4 e5 e6 
− η3 + d2 r2 + 4 + 7 + 10 − η 4 + e2 r2 + 4 + 7 + 10 + 13 ,
r r r r r r r r r
(184)
2
with η = σλ ≡ 21 Λ̃Λ2 being the small perturbation parameter. At zeroth order we have a1 = −2GM , a2 = λ. On the
other hand, b1 , c1 , d1 and e1 are constants of integration while the rest of the constants have dependence in these
first ones.
Clearly, for r  2GM , the asymptotic solution is the de Sitter spacetime, as expected,
  1
ds2 = 1 + Λ̃p r2 dt2 +   + r2 dΩ2 , (185)
D

1 + Λ̃p r2

with Λ̃p = a2 − b2 η − c2 η 2 − d2 η 3 − e2 η 4 .

C. Cosmology

The details of this section can be found in [135]. Again, let us start by considering only the subtle vanishing torsion
case. Considering a homogeneous and isotropic metric means that there exists a special spacetime foliation where each
spatial section is maximally symmetric50 . Therefore, the metric must be of a Friedmann-Lemaı̂tre-Robertson-Walker
(FLRW) type [134]

dr2
 
ds2 = −dt2 + a2 (t) + r 2
dΩ2
, (186)
1 − kr2

50 A maximally symmetric space is a space that has the same number of symmetries as the Euclidean space. In other words, a space is
maximally symmetric if it has n(n + 1)/2 linearly independent Killing vectors for an n-dimensional space
45

where a(t) is the scale factor, dΩ2 is the solid angle and the constant k ∈ {−1, 0, 1} defines the curvature of the spatial
sections. As usually used in the literature, it is convenient the following variables
 2
ä ȧ k
l≡ , h≡ + 2. (187)
a a a

We also define, ρΛ ≡ χ−1 Λ2 and ρΛ̃ ≡ χ−1 Λ̃2 . Hence, substituting the metric (186) in the field equations51 (176) and
(177), it is found
3 χ
h2 − h + (ρ + ρΛ̃ ) = 0 ,
2χρΛ 3
3 2 χ
l − l − (ρ + 3p − 2ρΛ̃ ) = 0 ,
2χρΛ 6
1
∂t (l + h) = 0 . (188)
(χρΛ )2

These are the complete equations governing the evolution of the cosmological scenario. We will explore them in

FT
different energy scale sectors. We notice that the third equation in (188) is a novelty information. In typical theories
this equation is not present. It arises from the fact that we have a curvature squared term and the fact that we are
considering the first order formalism.

1. The Λ-CDM model

The infrared sector is characterized by a low energy regime for gravity, i.e. weak curvature regime. The characteristic
energy scale of the model is given by Λ2 or, equivalently, through ρΛ . Thus, we have to compare the linear terms h
and l with ρΛ . If h and l  ρΛ , then the quadratic terms in (188) become negligible. Effectively, one can reach the
RA
IR regime by taking the limit Λ2 → +∞ in (176) and (177). Hence52 ,
χ
h= (ρ + ρΛ̃ ), (189)
3
χ
l = − (ρ + 3p − 2ρΛ̃ ), (190)
6
which are exactly the Einstein field equations with a cosmological term. Therefore, the present stage of the Universe
is correctly described by the model.

2. High curvature regime


D

Let us now consider a strong curvature regime in which the curvature square term is comparable with the other
terms. In contrast to the IR regime, now the quadratic curvature becomes comparable to ρΛ and can no longer
be neglected in (188). In what follows we shall outline separately the evolution of a vacuum from a matter filled
spacetime.
• Vacuum model
The dynamical equations for a vacuum universe in the UV sector follow from (188) by setting ρ = p = 0,
3 2 χ
h − h + ρΛ̃ = 0 ,
2χρΛ 3
3 2 χ
l − l + ρΛ̃ = 0 ,
2χρΛ 3
∂t (l + h) = 0 . (191)

51 Perfect fluit energy-momentum tensor has to be added in (177).


52 See Sect. XI A for the argumentation to drop the third equation (188).
46

The first two equations above are algebraic relations for h and l respectively. In fact, both equations have the
same structure showing that h and l share the same spectrum. It is straightforward to identify the roots as

Λ2
 r 
2 ρΛ̃
(h, l) = ΛdS± = 1± 1−2 , (192)
3 ρΛ

Since both h and l are constant, the third equation in (191) is automatically satisfied. The Ricci scalar reads
R = 6(l + h) which for l = h gives R = 12 Λ2dS± and for l 6= h we have R = 4Λ2 . The evolution of the scale
factor can automatically be integrated from the equation h = Λ2dS± . The possible solutions read

cosh (ΛdS± t) ; k = −1 ,
1 
a(t) = 2 exp (ΛdS± t) ; k = 0 , (193)
ΛdS± 
sinh (Λ
dS± t) ; k=1,
q
where ΛdS± ≡ Λ2dS± . One can immediately recognize these solutions being three different foliations of a de
Sitter universe with a cosmological constant given by Λ2dS± .

FT
The deceleration parameter is defined as q = − äa 2 −1
= −1/ 1 − k Λ−2 −2

ȧ2 = −l(h − k/a ) dS± a , which is always
negative. Hence, an accelerated expanding phase is manifest.
The effective cosmological constant depends both on ρΛ̃ and ρΛ . In the limit Λ̃2 /Λ2  1, we can expand up to
first order and obtain
" !#
2 2
Λ Λ̃
Λ2dS± ≈ 1± 1− 2 . (194)
3 Λ

The root Λ2dS+ is approximately proportional to Λ2 while the root Λ2dS− is approximately proportional to
RA
Λ̃2 . Given the enormous value of Λ2 , the first root represents an universe with a violent de Sitter phase.
Therefore, this phase can be associated with an inflationary expansion. On the other hand, the second root
would correspond to a smooth accelerating phase similar to the late time expansion in the ΛCDM model. This
result may be a Hint that a general solution connecting the inflationary era with the ΛCDM model could be
developed.
• Perfect fluid model
In the second order formalism one has to provide an equation of state that in this case reduces to give a
functional dependence of the pressure in terms of the energy density, i.e. p = p(ρ). In this case, the equation of
state of the fluid is of major importance to completely specific the dynamic system. In contrast, the first order
gravity analyzed in this paper has an extra set of dynamical field equations (188). They are a consequence of the
D

dynamically independent character of the spin connection field. In principle we have three variables to determine
the evolution, namely, the scale factor, the energy density and the pressure. However, with the thermodynamic
equation of state we would have four equation for three variables and the system would be overdetermined.
Indeed, the set of equations (188) is sufficient to establish the time evolution of all the variables. A possible way
to reconcile this situation with a thermodynamic description of matter is to interpretation the UV sector as a
regime where the gravitational field does not distinguish the nature of the matter fields, i.e. all perfect fluids
gravitate in the same manner in the UV regime. For the two thermodynamic quantities it is found (see fig. 4)

9ξ0 2 −8 R20
 
3ξ0 (4χρΛ − R0 ) −4 R0
ρ= a − a − ρ Λ̃ − + , (195)
4χ2 ρΛ 2χ2 ρΛ 4χ 32χ2 ρΛ
 
4χ − R0
p=−ρ+ ξ0 a−4 , (196)
χ2 ρΛ

The particular case ξ0 = 0 freezes the value of the energy density and the pressure becomes p = −ρ. Therefore,
we have to assume the general case where ξ0 6= 0. Equations (195) and (196) show the universal behavior of
the pressure and energy density independent of the equation of state of the fluid. Indeed, if one assumes a
barotropic equation of state p = ωρ with constant ω, equations (195) and (196) implies ω = −1. Thus, the only
consistent barotropic equation of state in the UV sector is p = −ρ. It is clear then that the energy density and
47

FIG. 4: Behaviour of pressure and energy density according to scale factor.

pressure given by (195) and (196) must be associated with a non-conservation of the energy-momentum tensor.
Indeed, equations (188) are consistent with

FT
ȧ ρ̇2
ρ̇ + 3 (ρ + p) + = 0, (197)
a 8(ȧ/a)(ρ + ρΛ̃ − ρΛ /2)
which verifies the non-conservation of the usual energy-momentum tensor. The first two terms are the usual
components while the last non-linear term comes from the high curvature corrections.
The scale factor can also be determined as
s
9k 2 − 3R0 ξ0 ∓αt 6k
a(t) = x0 e±αt + 2 e + , (198)
R0 x0 R0
RA
p
where α = R0 /3 and x0 > 0 is a constant of integration. The above solution has non-trivial R0 = 0 and
k = 0 limits.
Three different simple situations of (198) are displayed in Fig. 5. First, we take vanishing Ricci scalar but
nonzero spatial section, providing
q p
a(t) = ξ0 k − k(t ± |ξ0 |)2 . (199)
The ± sign within the square term does not change qualitatively the evolution. For k = 1 we have a big bang-big√
crunch solution with
√ an initial and a final singularity. The scale factor reaches its maximal value at tmax = ξ0
where a(tmax ) = ξ0 . The k = −1 has to disjoint branches. Hence, both branches have an initial singularity
D

and expands or it is a collapsing universe with a future singularity. For the plus√ sign the initial singularity is
located at t = 0 while the final singularity in the collapsing phase is at t = −2 ξ0 . For flat spatial section the

FIG. 5: Time evolution of the scale factor.

constant ξ0 must be positive definite and we have


q p
a(t) = ±2 ξ0 t, (200)
48

where the constant of integration was chosen to locate the singularity at t = 0. Once again we have a collapsing
phase for t < 0 that reaches the singularity and an expanding phase with initial singularity at t = 0.
Finally, let us only comment that the characteristic of the general solution (198) depends on the interplay of
the constants k, R0 and ξ0 . In particular, the evolution can describe a bounce if the second term has the same
sign as the first one, namely, if R0 < 3/ξ0 .

D. Dark stuff

Nowadays, dark matter and dark energy could be expected features from a good fundamental gravity theory. They
can have a completely different origin. However, the fact that dark matter only interacts only with gravity could be
a clue about its gravity origin.
Dark energy is not a problem if Λ e is at our disposal. Dark matter, on the other hand, is a bit more complicated.
There are two possible ways that the present approach could account for dark matter. The first one is to enlarge the
gauge group to a bigger group encoding the SO(5). A known example is the SL(5, R) group, as in [36, 40]. In this
case, the extra sector of the gauge group manifests itself in the geometric sector as a matter field which interacts only
with gravity. Moreover, the mass of this field is very large since it is proportional to Λ2 . This means that the range

FT
of this field is very very small. Hence, this field should very difficult to be detected, as expected from dark matter.
The other possibility is to remain with the SO(5) group and consider the contribution of the Gribov-Zwanziger fields.
This contribution is also small because these terms are proportional to ~2 , since it comes from 1-loop contributions.
Moreover, they would also interact only with gravity while a very small equivalence principle violation is expected
due to the non-covariance of the Faddeev-Popov operator. this last property is not discarded from a possible dark
matter behavior, see for instance [122–124].

Appendix A: Differential forms


RA
This appendix is not a rigorous exposition of exterior differential forms. We provide here a brief survey about the
main definitions and results only for operational matters. In other words, we provide here a kind of practical tutorial
about differential forms. For details we refer to the vast existing literature on the topic, see for instance [62, 125–128].

1. Exterior product and p-forms

For generality purposes, let us consider D-dimensional manifold M D as the starting scenario. For each differentiable
curve passing at a generic point x ∈ M D , there is a tangent vector field. The tangent space Tx (M D ) is the collection
of all tangent vectors at x. The differential operators {∂µ }, with µ ∈ {1, · · · , D}, are the components for which a
basis in Tx (M D ) is constructed. The dual space Tx∗ (M D ) is called cotangent space and its basis is defined by the
D

objects dxµ . Such duality is imposed through the inner product rule
(∂µ , dxν ) = δµν . (A1)

The exterior product is typically represented by the symbol ∧. For two components of the basis for Tx∗ (M D ), the
exterior product is defined as the anti-commuting product
dxµ ∧ dxν = dxµ ⊗ dxν − dxν ⊗ dxµ = −dxν ∧ dxµ . (A2)
In the case of µ = ν we have a null exterior product, obviously. It is easy to see that the exterior product is associative.
For the sake of simplicity, from now on, we omit the ∧ symbol from the notation.
An exterior differential form ω, or simply a p-form, is defined by
1
ω= ωµ µ ···µ dxµ1 dxµ2 · · · dxµp , (A3)
p! 1 2 p
where ωµ1 µ2 ···µp are covariant tensor components of order p 6 d. Due to (A2), these components are completely
anti-symmetric on their indexes. Moreover, the restriction p 6 d follows from the obvious fac
The exterior product between a p-form α and a q-form β results in a r-form γ such that r = p + q, which is
non-vanishing if r 6 d, given by
γ = αβ = (−1)pq βα , (A4)
49

where the factor ±1 comes from the anti-commuting nature of the exterior product: two generic forms will anti-
commute only if both forms are odd in their ranks.

2. Exterior derivative

The differential operator d, which is called exterior derivative, is defined by

d = ∂ν dxν . (A5)

It acts on p-form as follows


1 1
(∂ν ) dxν ωµ1 µ2 ···µp dxµ1 dxµ2 · · · dxµp = ∂ν ωµ1 µ2 ···µp dxν dxµ1 dxµ2 · · · dxµp .

dω = (A6)
p! p!
It is evident that the exterior derivative is a map from the space of p-forms to the space of (p + 1)-form.
The exterior derivative has two important properties. The first of them concerns its nilpotency, d2 = 0. The second
is about its application on a product of forms and the Leibniz rule. To illustrate this second point, let us take a

FT
p-form α and a q-form β. The action of the exterior derivative on the product is explicitly shown below,

d(αβ) = dαβ + (−1)p αdβ . (A7)

If the exterior derivative acts on a sum of forms, then

d(c1 α + c2 β) = c1 dα + c2 dβ , (A8)

where c1 and c2 are 0-forms.


RA
3. The Hodge dual operator

The Hodge duality consists in an operation that maps a p-form into a dual (D − p)-form. The Hodge dual operator
is a map as follows,

g µ ···µ
∗ dxµ1 dxµ2 · · · dxµp =  1 p µp+1 ···µd dxµp+1 · · · dxµD , (A9)
(D − p)!

where g = | det gµν |. Hence, the Hodge dual associated to a p-form ω is



g µ ···µ
∗ω = ωµ ···µ  1 p µp+1 ···µd dxµp+1 · · · dxµD . (A10)
p!(D − p)! 1 p
D

The Hodge duality is an isomorphism that maps the space of all p-forms Ep in the space of all D − p-forms ED−p as
?
Ep 7−−→ ED−p . (A11)

Since the dimension of the space p-form is given by


 
Dp D!
dim E = ≡ , (A12)
p p!(D − p)!
it is straightforward to see that
 
D
dim ED−p = , (A13)
D−p
i.e., the dimension of the dual space has the same dimension of the space of p-forms. The Hodge dual operator satisfies
the linear property,

∗ (c1 α + c2 β) = c1 ∗ α + c2 ∗ β , (A14)

where c1 and c2 are 0-forms. When the Hodge dual operator is twice employed, we obtain
50
(
(−1)p(D−p) if MD is Euclidean,
?? = .
(−1)p(D−p)+1 if MD is Minkowskian.
Hence, the double Hodge dual operation brings back the p-form to the original shape up to a signal. Another
interesting mathematical relation is about Levi-Civita tensor. In a D-dimensional Euclidean space the volume form
is related to that tensor as
1
dx1 · · · dxD = µ ···µ dxµ1 · · · dxµD ⇒ dxµ1 · · · dxµD = µ1 ···µD dx1 · · · dxD , (A15)
D! 1 D
while in a D-dimensional Minkowskian space we have
1
dx0 · · · dxD−1 = µ ···µ dxµ1 · · · dxµD ⇒ dxµ1 · · · dxµD = −µ1 ···µD dx0 · · · dxD−1 , (A16)
D! 1 D
since the full contraction
( between the Levi-Civita tensor is given as follow,
D! for Euclidean space
µ1 ···µD µ1 ···µD =
−D! for Minkowskian space .

FT
It is also possible to employ the vielbein to project the Levi-Civita tensor indexes,

eaµ11 eaµ22 · · · eaµkk a1 a2 ···ak µk+1 µk+2 ···µk+` = µ1 µ2 ···µk µk+1 µk+2 ···µk+` ,
eaµ11 eaµ22 · · · eaµkk a1 a2 ···ak µk+1 µk+2 ···µk+` = µ1 µ2 ···µk µk+1 µk+2 ···µk+` . (A17)

The inverse relations are easily found

eaµ11 eaµ22 · · · eaµkk µ1 µ2 ···µk ak+1 ak+2 ···ak+` = a1 a2 ···ak ak+1 ak+2 ···ak+` ,
eaµ11 eaµ22 · · · eaµkk µ1 µ2 ···µk ak+1 ak+2 ···ak+` = a1 a2 ···ak ak+1 ak+2 ···ak+` . (A18)
RA
4. Integration of differential forms

The integral of a p-form ω over an oriented region U ∈ Rp is defined as


Z Z Z
1 µ1 µp 1
ω= ωµ1 ···µp dx · · · dx = o(x) ωµ1 ···µp dx1 · · · dxp , (A19)
U U p! U p!

where o(x) = ±1 depending on the orientation of the coordinate basis. The result is, obviously, a (p − 1)-form.
An important property that uses the integration of p-forms is contained in the definition of the inner product. Let
α and β be generic p-forms. If α ∗ β is an r-form, then we define the inner product on a p-dimensional manifold M p
as
D

Z Z
(α, β) = α∗β = β∗α . (A20)
Md U

Finally, the Gauss-Stokes’ theorem in this formalism can be enunciated:


Z I
dω = ω. (A21)
U ∂U

Appendix B: Geometrical aspects of gauge theories

In this appendix we summarize some formal concepts about the geometrical structure of gauge theories. A parallel
with the gravity geometrical structure is also presented. Notions about topological spaces, maps and manifolds may
be required. Proofs are omitted. Standard references are [2, 62, 125]. See also [63] for a short discussion and [129]
for an advanced approach. For simplicity we employ differential form notation and set κ = 1.
51

1. Principal bundles and connections

We have learned about Lee groups at Sec. I C 1. Together with a differential manifold (spacetime, in this case), A
Lie group form the basic structure for gauge theories. Such structure is a sophisticated product of topological spaces
called principal bundles, which is a particular case of the more general concept of fibre bundles. Formally, a Fibre
bundle is a space B(G, M, F ), called total space, obeying the following conditions [62]
• B is a differential manifold;
• M , called the base space, is a differential manifold with an open covering {Ui };
• F , called fibre, is a differential manifold;
• G is a Lie group, called structure group, with left action on F ;
• ∃ a surjective map53 π : B 7−→ M , called projection map.
• ∃ a diffeomorphism54 φi : {Ui } × F 7−→ π −1 (Ui ) such that π ◦ φi (p, f ) = p where55 p ∈ M and f ∈ F .
Equivalently, φi (p, f ) : F 7−→ Fp is a diffeomorphism.

FT
• Whenever Ui ∩ Uj 6= ∅ we have tij (p) = φ−1
i (p, f ) ◦ φj (p, f ) : F 7−→ F ∈ G. Moreover, tij (p) must respect

– For p ∈ Ui , tii (p) = id;


– For p ∈ Ui ∩ Uj , tij (p) = t−1
ji (p);
– For p ∈ Ui ∩ Uj ∩ Uk , tij (p)tjk (p) = tik (p);
The inverse of the projection defines the fiber at p ∈ M , π −1 : (p) = Fp , which is a “copy” of F at p. The
diffeomorphism φi is called the local trivialization and is analogous to the concept of local Cartesian product. The
maps tij is a smooth map between φi and φj and are called transition functions.
RA
Further important definitions are:

• A section on B is a smooth map s : M 7−→ B π ◦ s = id. Thus, s(p) = π −1 (p) ∈ Fp . A global section is a

section defined independently of Ui while a local section is only defined in Ui .
• A fibre bundle is a trivial bundle if all transition functions can be taken as identity maps, i.e. a trivial bundle
is a direct (Cartesian) product M × F . Equivalently, a trivial bundle is a fibre bundle which admits a global
section. Nevertheless, a fibre bundle is also trivial if the base space M is contractible to a point.
A principal bundle is simply a fibre bundle where the fibre coincides with the structure group, F = G, and is
denoted by P (G, M ). Besides the left action of G on F = G, the right action of G on the fibre can be defined by
means of: Let u ∈ G and q = (p, fi ) ∈ P then φ−1 i (qu) = (p, fi u). Applying φi (q) from the left in this last expression,
D

we find qu = φi (p, fi u). The right action is map Fp 7−→ Fp and is possible because of the closure property of groups.
A connection on a principal bundle is the fibre bundle analogue of affine connections in Riemannian manifolds. It
allows to compare vector in the tangent space of P (G, M ), providing the notion of parallelism in the total space P .
To do this, we split the tangent space of P , Tq (P ), at a point q = (p, g) ∈ P into its vertical and horizontal subspaces,
namely Vq and Hq , respectively. The vertical space is tangent to a fiber Gp . As a consequence, as discussed in
Sec. I C 1, it is equivalent to the algebra of G. The horizontal space is a complement of Vq , which is not unique.
Hence, a connection is a unique split of Tq (P ) obeying
• Tq (P ) = Hq ⊕ Vq
• A vector field X ∈ P split as X = X H + X Q , where X H ∈ Hq and X V ∈ Vq ;
• At the same fibre, horizontal spaces are related by the right action of G, i.e. Rg : Hq 7−→ Hqg .

53 A surjective map between two topological spaces is a map whose image is entirely covered by the map.
54 A diffeomorphism between two topological spaces is a smooth and continuous map whose inverse is also continuous and smooth, i.e. a
diffeomorphism is a smooth homeomorphism.
55 Obviously, q = (p, f ) is a point of the total space B.
52

Thus, a connection is a set of conditions that picks only one specific Hq amongst all possible horizontal subspaces.
In accordance with the above definitions (see [62]), it is possible to define the connection one-form as an algebra-
valued one-form ω ∈ Tq∗ (P ) × AlgG, which is a projection of Tq (P ) in Vq . The corresponding local connection form
is the pullback of ω to Ui , i.e. Ai = σi∗ ω ∈ Ω1 (Ui ) × AlgG where σi∗ is the pullback from q to p associated to a local
section σi and Ω1 (U )i) is the space of one-forms in Ui ∈ M . To ω be uniquely defined on Ui ∩ Uj , it is required that

Ai 7−→ t−1 ijAi t−1 −1


ij + tij dtij , (B1)

where d is the exterior derivative.


To make contact with the known concept of affine connections, we have to define the horizontal lift of a curve:
Given two curves c : [0, 1] 7−→ M and ec : [0, 1] 7−→ P , ec is a horizontal lift of c if π ◦ e
c 7−→ c and the tangent vector of
c belongs to Hec . It can be shown [62] that there exists a map π −1 (c(0)) 7−→ π −1 (c(1)) while q0 7−→ q1 which defines
e
the parallel transport of q0 through the curve e c in such a way that
 Z 
q1 = σi (c(1))O exp − Aiµ dxµ , (B2)
c

where σi (c(1)) is the local section at c(1) ∈ Ui , O is the path ordering operator and t ∈ [0, 1] is the parametrization

FT
of the curve. Eq. (B2) clearly states that the components of Ai are responsible for the parallel transport along c.
The covariant derivative rises in a natural way as the horizontal exterior derivative in P [62, 63], namely dP = D.
In Ui , the covariant derivative components are

D i = d + Ai . (B3)

The commutation between covariant derivatives measures the commutativity of the parallel transport of vector in P
along a curves e
c1 and e
c2 which have the same starting point e c2 (0). In fact, D2 = Dω leads to components of
c1 (0) = e
the form
RA
Fi = dAi + Ai Ai . (B4)

The basic structure of the gauge theories we are considering in this lectures is the principal bundle P (G, R4 ) where
the base space is the four-dimensional Euclidean spacetime, G is the gauge group. In a sense, P (GR4 ) is a formal
manner to localize the rigid Lie group G by making a copy of it at each point of the spacetime. Moreover, since R4 can
be contracted down to a point, P (G, R4 ) is a trivial bundle. The local connection form in P (G, R4 ) is then recognized
as the gauge field and (B1) is the corresponding gauge transformation in the notation of fibre bundles. Thus, the
gauge field is a natural geometrical quantity in the abstract space P (G, R4 ). The covariant derivative components
(B3) are recognized as the covariant derivative in gauge theories (32) while the curvature components (B4) is directly
identified with the field strength (36). Hence, a pure geometrical scenario is set to explore gauge theories. Given a
gauge invariant classical action for A and the respective field equations, the solution of the theory with appropriate
boundary conditions is equivalent to find the appropriate unique horizontal space Hq . For a quantum theory however,
D

we need another principal bundle which takes into account all possible gauge configurations, i.e. the set of all possible
horizontal spaces in P .

2. Universal bundles

The most suitable scenario for a quantum gauge theory is a universal bundle, denoted by U (G, A/G) where the
total space is the space of all possible gauge connections configurations U = A while the fibre and structure group
are the Lie group G. The base space is the space off all independent gauge connections A/G, i.e. the space of gauge
connections not related through gauge transformations. In other words, one defines equivalence classes in A and take
one representative A of each equivalence class, the set of all A is the base space and is called moduli space. The fibre
is constructed by π −1 which manifest itself through the gauge orbit (44),

Au = u−1 (A + d)u . (B5)

In this context, the gauge fixing is an attempt to define a global section in U , which is not possible because U is not
trivial since A/G is not contractible to a point. In fact, this is the mathematical kernel of Gribov ambiguities [15]
which states that is not possible to globally fix the gauge of a gauge theory, i.e. it is not possible to pick only one
representative of each equivalence class [A].
53

Besides Au and F u , there are two more natural structures in U , namely, the BRST operator and the Faddeev-Popov
ghost field. If one defines the functional exterior derivative in U by56
Z
δ
δ= d4 x δAuµ , (B6)
δAuµ

where δAuµ is the functional analogous of dxµ , it is possible to show that

δAu = −D(Au )(u−1 δu) + u−1 δAu . (B7)

Clearly, the first term is the derivative along the fibre G while the second term is the derivative along the base space
A/G. Let us consider the anti-commuting argument of the covariant derivative of the first term, c = u−1 δu, its
derivative reads57

δc = −u−1 δuu−1 δu = −cc . (B8)

We remark that, although c is an anti-commuting quantity, cc = c2 6= 0 because it is an algebra-valued quantity. If

FT
we consider δ along the fibre, we end up with

δ f ibre Auµ = −D(Au )c ,




δ f ibre c = −cc . (B9)

Thus, the BRST operator can be interpreted as the functional derivative in U along the fibre58 , s = δ . Moreover,

f ibre
c is recognized as the Faddeev-Popov ghost field. Geometrically, c is the Maurer-Cartan form on group space while
the second equation in (B9) is known as the Maurer-Cartan structure equation.
RA
3. Gravity geometrical structure

Gravity can also be described in terms of principal bundles and, hence, as a gauge theory. However, extra features
arise.
First, let us consider a differential manifold M 4 and the tangent spaces TX (M 4 ) at points X ∈ M 4 . The set of all
tangent spaces,
\
T = TX (M 4 ) , (B10)
X∈M 4

is called tangent bundle. The tangent bundle has M 4 as the base space, the fibre is TX (M 4 ) and the structure group
D

is the GL(4, R) group, characterizing the transition functions as local coordinate transformations. It should be clear
that a point q ∈ T is characterized by q = (X, v) where V is a vector v = v µ ∂µ while π : T 7−→ M 4 provides
π −1 (X) = TX (M 4 ). Analogously, the cotangent bundle T ∗ can be defined as the set of all cotangent spaces in M by
changing vectors by forms and the basis ∂µ by the dual basis dxµ .
The fundamental structure for gravity is the frame bundle or, equivalently, the coframe bundle. The coframe
bundle C is the equivalent of the principal bundle P (G, R4 ) of gauge theories. The coframe bundle is associated to
the cotangent bundle where the fibre at X ∈ M 4 is the set of all local coframes ea = eaµ dxµ that can be defined
in TX (M 4 ) and the base space is M 4 . An element of C is then q = (X, e). Since all frames are related through
GL(4, R4 ) transformations (right action), the fibre coincides with the GL(4, R4 ) group. Hence, since the transition
functions are the same as the cotangent bundle, tij ∈ GL(4, R4 ), the coframe bundle actually is a principal bundle
C = (GL(4, R4 ), M ). Since the right action of the group leaves the coordinates in p intact, it is associated with
gauge transformations while the left action is associated with coordinate transformations. Thus, C is the perfect
basic structure to define gravity. In fact, since GL(4, R4) is contractible to O(4), the reduced coframe bundle O =

56 Just like the usual exterior derivative, δ is an anti-commuting operator.


57 We have used δ)(u−1 u) = 0.
58 It is an easy exercise to show that, in algebra components, (B9) reduce to (58a-58b)
54

P (O(4), M 4 ) coincides with the mathematical structure of a Einstein-Cartan gravity. Moreover, the natural connection
in O is recognized as the spin-connection ω ab . In essence, gravity is a gauge theory where the gauge group is identified
with the local isometries of spacetime.
Although a background independent quantum version of any gravity theory appear to be inconsistent, its mathe-
matical structure is not, yet it is a bit more complicated than U , defined at the previous section... ♠

Acknowledgments

We are thankful to the organizing committee for the opportunity in giving this lectures at the X Escola do Centro
Brasileiro de Pesquisas Fı́sicas, 13-24 July 2015. ADP acknowledges SISSA and the Theoretical Particle Physics
division for hospitality and support. The Conselho Nacional de Desenvolvimento Cientı́fico e Tecnológico59 (CNPq-
Brazil) and the Pró-Reitoria de Pesquisa, Pós-Graduação e Inovação (PROPPI-UFF) are acknowledged for financial
support.

FT
[1] V. A. Rubakov, “Classical theory of gauge fields,” Princeton, USA: Univ. Pr. (2002) 444 p.
[2] R. A. Bertlmann, “Anomalies in quantum field theory,” Oxford, UK: Clarendon (1996) 566 p. (International series of
monographs on physics: 91)
[3] S. Weinberg, “The Quantum theory of fields. Vol. 1: Foundations,” Cambridge, UK: Univ. Pr. (1995) 609 p
[4] S. Weinberg, “The quantum theory of fields. Vol. 2: Modern applications,” Cambridge, UK: Univ. Pr. (1996) 489 p
[5] D. Bailin and A. Love, Bristol, Uk: Hilger ( 1986) 348 P. ( Graduate Student Series In Physics)
[6] C. Itzykson, J. B. Zuber, “Quantum Field Theory,” New York, Usa: Mcgraw-hill (1980) 705 P.(International Series In
Pure and Applied Physics).
[7] C. N. Yang and R. L. Mills, Phys. Rev. 96, 191 (1954).
[8] G. ’t Hooft, Hackensack, USA: World Scientific (2005) 487 p
RA
[9] J. B. Hartle, An introduction to Einstein’s general relativity, San Francisco, USA: Addison-Wesley (2003) 582 p
[10] S. M. Carroll, Spacetime and geometry: An introduction to general relativity, San Francisco, USA: Addison-Wesley (2004)
513 p
[11] C. W. Misner, K. S. Thorne and J. A. Wheeler, Gravitation, San Francisco 1973, 1279p
[12]
[12] S. Weinberg, Gravitation and cosmology: principles and applications of the general theory of relativity, Wiley (1972) 688
p
[13] R. M. Wald, General Relativity, Chicago, Usa: Univ. Pr. ( 1984) 491p
[14] V. N. Gribov, Nucl. Phys. B139, 1 (1978).
[15] I. M. Singer, Commun. Math. Phys. 60, 7 (1978).
[16] M. B. Green, J. H. Schwarz and E. Witten, “Superstring Theory. Vol. 1: Introduction,” Cambridge, Uk: Univ. Pr. (
1987) ( Cambridge Monographs On Mathematical Physics)
[17] M. B. Green, J. H. Schwarz and E. Witten, “Superstring Theory. Vol. 2: Loop Amplitudes, Anomalies And Phenomenol-
D

ogy,” Cambridge, Uk: Univ. Pr. ( 1987) ( Cambridge Monographs On Mathematical Physics)
[18] J. Polchinski, “String theory. Vol. 1: An introduction to the bosonic string,” Cambridge, UK: Univ. Pr. (1998) 402 p
[19] J. Polchinski, “String theory. Vol. 2: Superstring theory and beyond,” Cambridge, UK: Univ. Pr. (1998) 531 p
[20] C. Rovelli, “Covariant loop gravity,” Lect. Notes Phys. 863, 57 (2013).
[21] R. Gambini and J. Pullin, “A first course in loop quantum gravity,” Oxford, UK: Univ. Pr. (2011) 183 p
[22] R. Percacci, “Asymptotic Safety,” In *Oriti, D. (ed.): Approaches to quantum gravity* 111-128 [arXiv:0709.3851 [hep-th]].
[23] R. Percacci, “A Short introduction to asymptotic safety,” arXiv:1110.6389 [hep-th].
[24] P. Horava, Phys. Rev. D 79, 084008 (2009) [arXiv:0901.3775 [hep-th]].
[25] P. Horava and C. M. Melby-Thompson, Phys. Rev. D 82, 064027 (2010) [arXiv:1007.2410 [hep-th]].
[26] D. Vernieri and T. P. Sotiriou, J. Phys. Conf. Ser. 453, 012022 (2013) [arXiv:1212.4402 [hep-th]].
[27]
[27] H. Steinacker, JHEP 0712, 049 (2007) [arXiv:0708.2426 [hep-th]].
[28] H. Steinacker, Class. Quant. Grav. 27, 133001 (2010) [arXiv:1003.4134 [hep-th]].
[29] C. Barcelo, M. Visser and S. Liberati, Int. J. Mod. Phys. D 10, 799 (2001) [gr-qc/0106002].
[30] T. Padmanabhan, AIP Conf. Proc. 939, 114 (2007) [arXiv:0706.1654 [gr-qc]].
[31] K. S. Stelle, P. C. West, Phys. Rev. D21, 1466 (1980).

59 RFS is a level PQ-2 researcher under the program Produtividade em Pesquisa, 308845/2012-9.
55

[32] F. W. Hehl, G. D. Kerlick and P. Von Der Heyde, Phys. Lett. B 63, 446 (1976).
[33] S. W. MacDowell, F. Mansouri, Phys. Rev. Lett. 38, 739 (1977).
[34] H. R. Pagels, Phys. Rev. D29, 1690 (1984).
[35] R. Tresguerres, Int. J. Geom. Meth. Mod. Phys. 5, 171-183 (2008). [arXiv:0804.1129 [gr-qc]].
[36] E. W. Mielke, “Einsteinian gravity from a spontaneously broken topological BF theory,” Phys. Lett. B688, 273-277
(2010).
[37] R. F. Sobreiro and V. J. Vasquez Otoya, Class. Quant. Grav. 24, 4937 (2007) [arXiv:0705.2200 [hep-th]].
[38] R. F. Sobreiro, V. J. V. Otoya, J. Phys. Conf. Ser. 283, 012032 (2011). [arXiv:1010.2946 [hep-th]].
[39] R. F. Sobreiro, A. A. Tomaz and V. J. V. Otoya, Eur. Phys. J. C 72, 1991 (2012) [arXiv:1109.0016 [hep-th]].
[40] T. S. Assimos, A. D. Pereira, T. R. S. Santos, R. F. Sobreiro, A. A. Tomaz and V. J. V. Otoya, arXiv:1305.1468 [hep-th].
[41] R. F. Sobreiro, A. A. Tomaz and V. J. V. Otoya, PoS ICMP 2012, 019 (2012) [arXiv:1210.8446 [hep-th]].
[42] R. F. Sobreiro, A. A. Tomaz and V. J. Vasquez Otoya, J. Phys. Conf. Ser. 453, 012014 (2013) [arXiv:1211.5993 [hep-th]].
[43] O. Piguet and S. P. Sorella, “Algebraic renormalization: Perturbative renormalization, symmetries and anomalies,” Lect.
Notes Phys. M 28, 1 (1995).
[44] D. Zwanziger, Nucl. Phys. B399, 477-513 (1993).
[45] R. F. Sobreiro, S. P. Sorella, Lectures given by S.P. Sorella at the 13th Jorge Andre Swieca Summer School on Particles and
Fields, Campos de Jordão, Brazil, 9-22 January 2005. “Introduction to the Gribov ambiguities in Euclidean Yang-Mills
theories,”
[46] D. Dudal, R. F. Sobreiro, S. P. Sorella, H. Verschelde, Phys. Rev. D72, 014016 (2005). [hep-th/0502183].

FT
[47] M. A. L. Capri, V. E. R. Lemes, R. F. Sobreiro, S. P. Sorella and R. Thibes, Phys. Rev. D 77, 105023 (2008)
[arXiv:0801.0566 [hep-th]].
[48] D. Dudal, S. P. Sorella, N. Vandersickel, [arXiv:1105.3371 [hep-th]].
[49] M. A. L. Capri, A. D. Pereira, R. F. Sobreiro and S. P. Sorella, “Non-perturbative treatment of the linear covariant
gauges by taking into account the Gribov copies,” arXiv:1505.05467 [hep-th].
[50] M. A. L. Capri, D. Dudal, D. Fiorentini, M. S. Guimaraes, I. F. Justo, A. D. Pereira, B. W. Mintz and L. F. Palhares
et al., “An exact nilpotent non-perturbative BRST symmetry for the Gribov-Zwanziger action in the linear covariant
gauge,” arXiv:1506.06995 [hep-th].
[51] J. Schwinger, L. L. Deraad Jr., K. A. Milton, W. Tsai and J. Norton, “Classical electrodynamics” ♠ ...
[52] A. O. Barut, “Electrodynamics and classical theory of fields and particles,” New York, USA: Dover (1980) 235 p
[53] Y. Aharonov and D. Bohm, Phys. Rev. 115, 485 (1959).
RA
[54] J. J. Sakurai and J. Napolitano, “Modern quantum physics,” Boston, USA: Addison-Wesley (2011) 550 p
[55] H. Georgi, “Lie algebras in particle physics,” Front. in Phys. Westview Pr. (1999) 320p
[56] R. Gilmore, “Lie groups, physics, and geometry: An introduction for physicists, engineers and chemists,” Cambridge,
UK: Univ. Pr. (2008) 319 p
[57] P. Ramond, “Group theory: A physicist’s survey,” Cambridge, UK: Univ. Pr. (2010) 310 p
[58] L. D. Faddeev and V. N. Popov, Phys. Lett. B 25, 29 (1967).
[59] R. F. Streater and A. S. Wightman, “PCT, spin and statistics, and all that,” Princeton, USA: Princeton Univ. Pr. (2000)
207 p.
[60] C. Becchi, A. Rouet and R. Stora, Phys. Lett. B 52, 344 (1974).
[61] I. V. Tyutin, “Gauge Invariance in Field Theory and Statistical Physics in Operator Formalism,” Lebedev Institute
preprint (Unpublished), 75-39, 1975. arXiv:0812.0580 [hep-th].
[62] M. Nakahara, Geometry, topology and physics, Bristol, UK: Hilger (1990) 505 p. (Graduate student series in physics).
[63] M. Daniel and C. M. Viallet, Rev. Mod. Phys. 52, 175 (1980).
D

[64] T. Kugo and I. Ojima, Phys. Lett. B 73, 459 (1978).


[65] T. Kugo and I. Ojima, Prog. Theor. Phys. 60, 1869 (1978).
[66] D. J. Gross and F. Wilczek, Phys. Rev. Lett. 30, 1343 (1973).
[67] H. D. Politzer, Phys. Rev. Lett. 30, 1346 (1973).
[68] J. C. Collins, “Renormalization. An Introduction To Renormalization, The Renormalization Group, And The Operator
Product Expansion,” Cambridge, Uk: Univ. Pr. ( 1984) 380p
[69] R. K. Ellis, W. J. Stirling and B. R. Webber, “QCD and collider physics,” Camb. Monogr. Part. Phys. Nucl. Phys.
Cosmol. 8, 1 (1996).
[70] D. Zwanziger, Nucl. Phys. B 323, 513 (1989).
[71] D. Dudal, J. A. Gracey, S. P. Sorella, N. Vandersickel and H. Verschelde, Phys. Rev. D 78, 065047 (2008) [arXiv:0806.4348
[hep-th]].
[72] N. Vandersickel and D. Zwanziger, Phys. Rept. 520, 175 (2012) [arXiv:1202.1491 [hep-th]].
[73] N. Vandersickel, arXiv:1104.1315 [hep-th].
[74] D. Zwanziger, Nucl. Phys. B 209, 336 (1982).
[75] G. Dell’Antonio and D. Zwanziger, Commun. Math. Phys. 138, 291 (1991).
[76] P. van Baal, Nucl. Phys. B 369, 259 (1992).
[77] A. J. Gomez, M. S. Guimaraes, R. F. Sobreiro and S. P. Sorella, Phys. Lett. B 683, 217 (2010) [arXiv:0910.3596 [hep-th]].
[78] M. A. L. Capri, D. Dudal, M. S. Guimaraes, L. F. Palhares and S. P. Sorella, Phys. Lett. B 719, 448 (2013)
[arXiv:1212.2419 [hep-th]].
[79] A. Cucchieri and T. Mendes, Phys. Rev. Lett. 100, 241601 (2008) [arXiv:0712.3517 [hep-lat]].
[80] A. Cucchieri and T. Mendes, Phys. Rev. D 78, 094503 (2008) [arXiv:0804.2371 [hep-lat]].
56

[81] A. Cucchieri, D. Dudal, T. Mendes and N. Vandersickel, Phys. Rev. D 85, 094513 (2012) [arXiv:1111.2327 [hep-lat]].
[82] A. D. Pereira and R. F. Sobreiro, Eur. Phys. J. C 73, 2584 (2013) [arXiv:1308.4159 [hep-th]].
[83] D. Dudal, M. S. Guimaraes and S. P. Sorella, Phys. Rev. Lett. 106, 062003 (2011) [arXiv:1010.3638 [hep-th]].
[84] D. Dudal, M. S. Guimaraes and S. P. Sorella, Phys. Lett. B 732, 247 (2014) [arXiv:1310.2016 [hep-ph]].
[85] F. Canfora and L. Rosa, Phys. Rev. D 88, 045025 (2013) [arXiv:1308.1582 [hep-th]].
[86] M. A. L. Capri, M. S. Guimaraes, I. F. Justo, L. F. Palhares and S. P. Sorella, Phys. Rev. D 90, no. 8, 085010 (2014)
[arXiv:1408.3597 [hep-th]].
[87] A. D. Pereira, Jr. and R. F. Sobreiro, Eur. Phys. J. C 74, no. 8, 2984 (2014) [arXiv:1402.3477 [hep-th]].
[88] M. A. L. Capri, A. J. Gomez, M. S. Guimaraes, V. E. R. Lemes and S. P. Sorella, J. Phys. A 43, 245402 (2010)
[arXiv:1002.1659 [hep-th]].
[89] A. Cucchieri, D. Dudal, T. Mendes and N. Vandersickel, Phys. Rev. D 90, no. 5, 051501 (2014) [arXiv:1405.1547 [hep-lat]].
[90] R. Utiyama, Phys. Rev. 101, 1597 (1956).
[91] T. W. B. Kibble, J. Math. Phys. 2, 212 (1961).
[92] D. W. Sciama, Rev. Mod. Phys. 36, 463-469 (1964).
[93] A. Mardones and J. Zanelli, Class. Quant. Grav. 8, 1545 (1991).
[94] J. Zanelli, “Lecture notes on Chern-Simons (super-)gravities. Second edition (February 2008),” arXiv:hep-th/0502193.
[95] V. De Sabbata and M. Gasperini, Introduction To Gravity, Singapore, Singapore: World Scientific ( 1985) 346p
[96] R. M. Wald, Space, time, and gravity: The Theory of the big bang and black holes, Chicago, USA: Univ. Pr. (1992) 153 p
[97] L. Ryder, Introduction to general relativity, Cambridge, UK: Cambridge Univ. Pr. (2009) 441 p

FT
[98] T. Padmanabhan, Gravitation: Foundations and frontiers, Cambridge, UK: Cambridge Univ. Pr. (2010) 700 p
[99] A. Einstein, The Foundation of the General Theory of Relativity, Annalen Phys. 49, 769 (1916) [Annalen Phys. 14, 517
(2005)].
[100] Deduzione invariantiva delle equazioni gravitazionali dal principio di Hamilton Rend. Circ. Mat. Palermo 43: 203212.
[101] E. Cartan, Sur les varits connexion affine et la thorie de la relativit gnralise. (premire partie), Annales Sci. Ecole Norm.
Sup. 40, 325 (1923).
[102] E. Cartan, Sur les varits connexion affine et la thorie de la relativit gnralise. (premire partie) (Suite), Annales Sci. Ecole
Norm. Sup. 41, 1 (1924).
[103] D. Lovelock, The Einstein tensor and its generalizations, J. Math. Phys. 12, 498 (1971).
[104] A. Achucarro and P. K. Townsend, A Chern-Simons Action for Three-Dimensional anti-De Sitter Supergravity Theories,
Phys. Lett. B 180, 89 (1986).
RA
[105] H. T. Nieh and C. N. Yang, A torsional topological invariant, Int. J. Mod. Phys. A 22, 5237 (2007).
[106] J. Doe, Colocar refs aqui!!!
[107] R. F. Sobreiro, Fiber bundles, gauge theories and gravity in Quantum Gravity ed. R. F. Sobreiro (Croatia: InTech) ISBN
978-953-51-0089-8.
[108] E. Witten, Nucl. Phys. B 311, 46 (1988).
[109] E. Inönü, E. P. Wigner, Proc. Nat. Acad. Sci. 39, 510-524 (1953).
[110] S. Weinberg and E. Witten, Phys. Lett. B 96, 59 (1980).
[111] L. Perivolaropoulos, Phys. Rev. D 77, 107301 (2008) [arXiv:0802.1531 [astro-ph]].
[112] T. Padmanabhan, arXiv:1210.4174 [hep-th].
[113] D. J. Toms, Phys. Lett. B 126, 37 (1983).
[114] I. L. Buchbinder, Theor. Math. Phys. 61, 1215 (1984) [Teor. Mat. Fiz. 61, 393 (1984)].
[115] L. Parker and D. J. Toms, Phys. Rev. D 32, 1409 (1985).
[116] I. L. Shapiro, J. Sola, On the possible running of the cosmological ’constant’, Phys. Lett. B682, 105-113 (2009).
D

[arXiv:0910.4925 [hep-th]].
[117] R. F. Sobreiro and A. A. Tomaz, “Predicting Planck scale and Newton and cosmological constants from a Yang-Mills
gauge theory: One and Two-loop estimates,” Work in progress (2015).
[118] A. Maas, Phys. Rept. 524, 203 (2013) [arXiv:1106.3942 [hep-ph]].
[119] A. Cucchieri and T. Mendes, Braz. J. Phys. 37, 484 (2007) [hep-ph/0605224].
[120] A. Cucchieri, T. Mendes, Landau-gauge propagators in Yang-Mills theories at beta = 0: Massive solution versus conformal
scaling, Phys. Rev. D81, 016005 (2010). [arXiv:0904.4033 [hep-lat]].
[121] S. Weinberg, Rev. Mod. Phys. 61, 1 (1989).
[122] O. Bertolami, F. Gil Pedro and M. Le Delliou, Dark Energy-Dark Matter Interaction and the Violation of the Equivalence
Principle from the Abell Cluster A586, Phys. Lett. B 654, 165 (2007) [astro-ph/0703462 [ASTRO-PH]].
citeCarroll:2008ub
[123] S. M. Carroll, S. Mantry, M. J. Ramsey-Musolf and C. W. Stubbs, Dark-Matter-Induced Weak Equivalence Principle
Violation, Phys. Rev. Lett. 103, 011301 (2009) [arXiv:0807.4363 [hep-ph]].
[124] J. A. Keselman, A. Nusser and P. J. E. Peebles, Cosmology with Equivalence Principle Breaking in the Dark Sector, Phys.
Rev. D 81, 063521 (2010) [arXiv:0912.4177 [astro-ph.CO]].
[125] C. Nash and S. Sen, Topology And Geometry For Physicists, London, Uk: Academic (1983) 311p.
[126] T. Frankel, The geometry of physics: An introduction 2nd Ed., Cambridge Univ. Pr. (2004) 698 p
[127] J. M. F. Bassalo and M. S. D. Cattani, Cálculo exterior, Ed. Livraria da Fı́sica (2009) 191 p
[128] H. Flanders, Differential Forms with Applications to the Physical Sciences, Dover Publications (2012) 224 p
[129] S. Kobayashi and K. Nomizu, Foundations of Differential Geometry - Vol. 1, New York, USA: John Wiley & Sons (1963).
[130] G. ’t Hooft, M. J. G. Veltman, Annales Poincare Phys. Theor. A20, 69-94
57

[131] S. Deser, P. van Nieuwenhuizen, Phys. Rev. D10, 401 (1974).


[132] S. Deser, P. van Nieuwenhuizen, Phys. Rev. D10, 411 (1974).
[133] I. L. Buchbinder, S. D. Odintsov and I. L. Shapiro, “Effective action in quantum gravity,” Bristol, UK: IOP (1992) 413 p
[134] S. Weinberg, Cosmology, Oxford, UK: Oxford Univ. Pr. (2008) 593 p
[135] F. T. Falciano, G. S. A. Sadovski, R. F. Sobreiro and A. A. Tomaz, “Cosmology of an SO(m, n) gauge induced gravity”,
Work in progress.
[136] A. Ashtekar, M. Reuter and C. Rovelli, “From General Relativity to Quantum Gravity,” arXiv:1408.4336 [gr-qc].

FT
RA
D

You might also like