LQG
LQG
LQG
History
Main article: History of loop quantum gravity
In 1986, Abhay Ashtekar reformulated Einstein's general relativity in a language
closer to that of the rest of fundamental physics, specifically Yang–Mills theory.
[2] Shortly after, Ted Jacobson and Lee Smolin realized that the formal equation of
quantum gravity, called the Wheeler–DeWitt equation, admitted solutions labelled by
loops when rewritten in the new Ashtekar variables. Carlo Rovelli and Smolin
defined a nonperturbative and background-independent quantum theory of gravity in
terms of these loop solutions. Jorge Pullin and Jerzy Lewandowski understood that
the intersections of the loops are essential for the consistency of the theory, and
the theory should be formulated in terms of intersecting loops, or graphs.
In 1994, Rovelli and Smolin showed that the quantum operators of the theory
associated to area and volume have a discrete spectrum. That is, geometry is
quantized. This result defines an explicit basis of states of quantum geometry,
which turned out to be labelled by Roger Penrose's spin networks, which are graphs
labelled by spins.
The canonical version of the dynamics was established by Thomas Thiemann, who
defined an anomaly-free Hamiltonian operator and showed the existence of a
mathematically consistent background-independent theory. The covariant, or "spin
foam", version of the dynamics was developed jointly over several decades by
research groups in France, Canada, UK, Poland, and Germany. It was completed in
2008, leading to the definition of a family of transition amplitudes, which in the
classical limit can be shown to be related to a family of truncations of general
relativity.[3] The finiteness of these amplitudes was proven in 2011.[4][5] It
requires the existence of a positive cosmological constant, which is consistent
with observed acceleration in the expansion of the Universe.
Background independence
LQG is formally background independent, meaning the equations of LQG are not
embedded in, or dependent on, space and time (except for its invariant topology).
Instead, they are expected to give rise to space and time at distances which are 10
times the Planck length. The issue of background independence in LQG still has some
unresolved subtleties. For example, some derivations require a fixed choice of the
topology, while any consistent quantum theory of gravity should include topology
change as a dynamical process.[citation needed]
Space-time as a "container" over which physics takes place has no objective
physical meaning and instead the gravitational interaction is represented as just
one of the fields forming the world. This is known as the relationalist
interpretation of space-time. In LQG this aspect of general relativity is taken
seriously and this symmetry is preserved by requiring that the physical states
remain invariant under the generators of diffeomorphisms. The interpretation of
this condition is well understood for purely spatial diffeomorphisms. However, the
understanding of diffeomorphisms involving time (the Hamiltonian constraint) is
more subtle because it is related to dynamics and the so-called "slave nature of
time" in general relativity.[6] A generally accepted calculational framework to
account for this constraint has yet to be found.[7][8] A plausible candidate for
the quantum Hamiltonian constraint is the operator introduced by Thiemann.[9]
{
�
�
,
�
}
�
�
=
�
�
=
�
=
0
=
{
�
�
,
�
}
�
�
=
�
�
=
�
=
0
=
{
�
,
�
}
�
�
=
�
�
=
�
=
0
=
0
,
{\displaystyle \{G_{j},O\}_{G_{j}=C_{a}=H=0}=\{C_{a},O\}_{G_{j}=C_{a}=H=0}=\
{H,O\}_{G_{j}=C_{a}=H=0}=0,}
that is, they are quantities defined on the constraint surface that are invariant
under the gauge transformations of the theory.
det
(
�
)
�
�
�
=
�
~
�
�
�
~
�
�
�
�
�
.
{\displaystyle \det(q)q^{ab}={\tilde {E}}_{i}^{a}{\tilde {E}}_{j}^{b}\delta ^{ij}.}
(where
�
�
�
\delta ^{ij} is the flat space metric, and the above equation expresses that
�
�
�
q^{ab}, when written in terms of the basis
�
�
�
E_{i}^{a}, is locally flat). (Formulating general relativity with triads instead of
metrics was not new.) The densitized triads are not unique, and in fact one can
perform a local in space rotation with respect to the internal indices
�
i. The canonically conjugate variable is related to the extrinsic curvature by
�
�
�
=
�
�
�
�
~
�
�
/
det
(
�
)
{\textstyle K_{a}^{i}=K_{ab}{\tilde {E}}^{ai}/{\sqrt {\det(q)}}}. But problems
similar to using the metric formulation arise when one tries to quantize the
theory. Ashtekar's new insight was to introduce a new configuration variable,
�
�
�
=
Γ
�
�
−
�
�
�
�
A_{a}^{i}=\Gamma _{a}^{i}-iK_{a}^{i}
The expressions for the constraints in Ashtekar variables; Gauss's theorem, the
spatial diffeomorphism constraint and the (densitized) Hamiltonian constraint then
read:
�
�
=
�
�
�
~
�
�
=
0
G^{i}={\mathcal {D}}_{a}{\tilde {E}}_{i}^{a}=0
�
�
=
�
~
�
�
�
�
�
�
−
�
�
�
(
�
�
�
~
�
�
)
=
�
�
−
�
�
�
�
�
=
0
,
{\displaystyle C_{a}={\tilde {E}}_{i}^{b}F_{ab}^{i}-A_{a}^{i}({\mathcal {D}}_{b}{\
tilde {E}}_{i}^{b})=V_{a}-A_{a}^{i}G^{i}=0,}
�
~
=
�
�
�
�
�
~
�
�
�
~
�
�
�
�
�
�
=
0
{\displaystyle {\tilde {H}}=\epsilon _{ijk}{\tilde {E}}_{i}^{a}{\tilde
{E}}_{j}^{b}F_{ab}^{k}=0}
respectively, where
�
�
�
�
F_{ab}^{i} is the field strength tensor of the connection
�
�
�
A_{a}^{i} and where
�
�
V_{a} is referred to as the vector constraint. The above-mentioned local in space
rotational invariance is the original of the
SU
(
2
)
\operatorname {SU} (2) gauge invariance here expressed by Gauss's theorem. Note
that these constraints are polynomial in the fundamental variables, unlike the
constraints in the metric formulation. This dramatic simplification seemed to open
up the way to quantizing the constraints. (See the article Self-dual Palatini
action for a derivation of Ashtekar's formalism).
With Ashtekar's new variables, given the configuration variable
�
�
�
A_{a}^{i}, it is natural to consider wavefunctions
Ψ
(
�
�
�
)
\Psi (A_{a}^{i}). This is the connection representation. It is analogous to
ordinary quantum mechanics with configuration variable
�
q and wavefunctions
�
(
�
)
\psi (q). The configuration variable gets promoted to a quantum operator via:
�
^
�
�
Ψ
(
�
)
=
�
�
�
Ψ
(
�
)
,
{\displaystyle {\hat {A}}_{a}^{i}\Psi (A)=A_{a}^{i}\Psi (A),}
(analogous to
�
^
�
(
�
)
=
�
�
(
�
)
{\hat {q}}\psi (q)=q\psi (q)) and the triads are (functional) derivatives,
�
�
�
~
^
Ψ
(
�
)
=
−
�
�
Ψ
(
�
)
�
�
�
�
.
{\displaystyle {\hat {\tilde {E_{i}^{a}}}}\Psi (A)=-i{\delta \Psi (A) \over \delta
A_{a}^{i}}.}
(analogous to
�
^
�
(
�
)
=
−
�
ℏ
�
�
(
�
)
/
�
�
{\hat {p}}\psi (q)=-i\hbar d\psi (q)/dq). In passing over to the quantum theory the
constraints become operators on a kinematic Hilbert space (the unconstrained
SU
(
2
)
\operatorname {SU} (2) Yang–Mills Hilbert space). Note that different ordering of
the
�
A's and
�
~{\tilde {E}}'s when replacing the
�
~{\tilde {E}}'s with derivatives give rise to different operators – the choice made
is called the factor ordering and should be chosen via physical reasoning. Formally
they read
�
^
�
|
�
⟩
=
0
{\hat {G}}_{j}\vert \psi \rangle =0
�
^
�
|
�
⟩
=
0
{\hat {C}}_{a}\vert \psi \rangle =0
�
~
^
|
�
⟩
=
0.
{\displaystyle {\hat {\tilde {H}}}\vert \psi \rangle =0.}
There are still problems in properly defining all these equations and solving them.
For example, the Hamiltonian constraint Ashtekar worked with was the densitized
version instead of the original Hamiltonian, that is, he worked with
�
~
=
det
(
�
)
�
{\textstyle {\tilde {H}}={\sqrt {\det(q)}}H}. There were serious difficulties in
promoting this quantity to a quantum operator. Moreover, although Ashtekar
variables had the virtue of simplifying the Hamiltonian, they are complex. When one
quantizes the theory, it is difficult to ensure that one recovers real general
relativity as opposed to complex general relativity.
{
�
(
�
)
,
�
�
�
}
=
∂
�
�
�
+
�
�
�
�
�
�
�
�
�
�
=
(
�
�
�
)
�
.
{\displaystyle \{G(\lambda ),A_{a}^{i}\}=\partial _{a}\lambda ^{i}+g\epsilon
^{ijk}A_{a}^{j}\lambda ^{k}=(D_{a}\lambda )^{i}.}
The quantum Gauss' law reads
�
^
�
Ψ
(
�
)
=
−
�
�
�
�
�
Ψ
[
�
]
�
�
�
�
=
0.
{\hat {G}}_{j}\Psi (A)=-iD_{a}{\delta \lambda \Psi [A] \over \delta A_{a}^{j}}=0.
If one smears the quantum Gauss' law and study its action on the quantum state one
finds that the action of the constraint on the quantum state is equivalent to
shifting the argument of
Ψ\Psi by an infinitesimal (in the sense of the parameter
�\lambda small) gauge transformation,
[
1
+
∫
�
3
�
�
�
(
�
)
�
^
�
]
Ψ
(
�
)
=
Ψ
[
�
+
�
�
]
=
Ψ
[
�
]
,
{\displaystyle \left[1+\int d^{3}x\lambda ^{j}(x){\hat {G}}_{j}\right]\Psi (A)=\Psi
[A+D\lambda ]=\Psi [A],}
and the last identity comes from the fact that the constraint annihilates the
state. So the constraint, as a quantum operator, is imposing the same symmetry that
its vanishing imposed classically: it is telling us that the functions
Ψ
[
�
]
\Psi [A] have to be gauge invariant functions of the connection. The same idea is
true for the other constraints.
Therefore, the two step process in the classical theory of solving the constraints
�
�
=
0
C_{I}=0 (equivalent to solving the admissibility conditions for the initial data)
and looking for the gauge orbits (solving the 'evolution' equations) is replaced by
a one step process in the quantum theory, namely looking for solutions
Ψ\Psi of the quantum equations
�
^
�
Ψ
=
0
{\hat {C}}_{I}\Psi =0. This is because it obviously solves the constraint at the
quantum level and it simultaneously looks for states that are gauge invariant
because
�
^
�
{\hat {C}}_{I} is the quantum generator of gauge transformations (gauge invariant
functions are constant along the gauge orbits and thus characterize them).[11]
Recall that, at the classical level, solving the admissibility conditions and
evolution equations was equivalent to solving all of Einstein's field equations,
this underlines the central role of the quantum constraint equations in canonical
quantum gravity.
LQG includes the concept of a holonomy. A holonomy is a measure of how much the
initial and final values of a spinor or vector differ after parallel transport
around a closed loop; it is denoted
ℎ
�
[
�
]
h_{\gamma }[A].
Knowledge of the holonomies is equivalent to knowledge of the connection, up to
gauge equivalence. Holonomies can also be associated with an edge; under a Gauss
Law these transform as
(
ℎ
�
′
)
�
�
=
�
�
�
−
1
(
�
)
(
ℎ
�
)
�
�
�
�
�
(
�
)
.
(h'_{e})_{{\alpha \beta }}=U_{{\alpha \gamma }}^{{-1}}(x)(h_{e})_{{\gamma \
sigma }}U_{{\sigma \beta }}(y).
For a closed loop
�
=
�
x=y and assuming
�
=
�\alpha =\beta , yields
(
ℎ
�
′
)
�
�
=
�
�
�
−
1
(
�
)
(
ℎ
�
)
�
�
�
�
�
(
�
)
=
[
�
�
�
(
�
)
�
�
�
−
1
(
�
)
]
(
ℎ
�
)
�
�
=
�
�
�
(
ℎ
�
)
�
�
=
(
ℎ
�
)
�
�(h'_{e})_{\alpha \alpha }=U_{\alpha \gamma }^{-1}(x)(h_{e})_{\gamma \sigma }U_{\
sigma \alpha }(x)=[U_{\sigma \alpha }(x)U_{\alpha \gamma }^{-1}(x)](h_{e})_{\
gamma \sigma }=\delta _{\sigma \gamma }(h_{e})_{\gamma \sigma }=(h_{e})_{\gamma \
gamma }
or
Tr
ℎ
�
′
=
Tr
ℎ
�
.
\operatorname {Tr} h'_{\gamma }=\operatorname {Tr} h_{\gamma }.
The trace of an holonomy around a closed loop is written
�
�
[
�
]
W_{\gamma }[A]
and is called a Wilson loop. Thus Wilson loops are gauge invariant. The explicit
form of the Holonomy is
ℎ
�
[
�
]
=
�
exp
{
−
∫
�
0
�
1
�
�
�
˙
�
�
�
�
(
�
(
�
)
)
�
�
}
{\displaystyle h_{\gamma }[A]={\mathcal {P}}\exp \left\{-\int _{\gamma _{0}}^{\
gamma _{1}}ds{\dot {\gamma }}^{a}A_{a}^{i}(\gamma (s))T_{i}\right\}}
where
�\gamma is the curve along which the holonomy is evaluated, and
�
s is a parameter along the curve,
�
{\mathcal {P}} denotes path ordering meaning factors for smaller values of
�
s appear to the left, and
�
�
T_{i} are matrices that satisfy the
SU
(
2
)
\operatorname {SU} (2) algebra
[
�
�
,
�
�
]
=
2
�
�
�
�
�
�
�
.
{\displaystyle [T^{i},T^{j}]=2i\epsilon ^{ijk}T_{k}.}
The Pauli matrices satisfy the above relation. It turns out that there are
infinitely many more examples of sets of matrices that satisfy these relations,
where each set comprises
(
�
+
1
)
×
(
�
+
1
)
(N+1)\times (N+1) matrices with
�
=
1
,
2
,
3
,
…N=1,2,3,\dots , and where none of these can be thought to 'decompose' into two or
more examples of lower dimension. They are called different irreducible
representations of the
SU
(
2
)
\operatorname {SU} (2) algebra. The most fundamental representation being the Pauli
matrices. The holonomy is labelled by a half integer
�
/
2
N/2 according to the irreducible representation used.
The use of Wilson loops explicitly solves the Gauss gauge constraint. Loop
representation is required to handle the spatial diffeomorphism constraint. With
Wilson loops as a basis, any Gauss gauge invariant function expands as,
Ψ
[
�
]
=
∑
�
Ψ
[
�
]
�
�
[
�
]
.
{\displaystyle \Psi [A]=\sum _{\gamma }\Psi [\gamma ]W_{\gamma }[A].}
This is called the loop transform and is analogous to the momentum representation
in quantum mechanics (see Position and momentum space). The QM representation has a
basis of states
exp
(
�
�
�
)
\exp(ikx) labelled by a number
�
k and expands as
�
[
�
]
=
∫
�
�
�
(
�
)
exp
(
�
�
�
)
.
\psi [x]=\int dk\psi (k)\exp(ikx).
and works with the coefficients of the expansion
�
(
�
)
.
{\displaystyle \psi (k).}
Ψ
[
�
]
=
∫
[
�
�
]
Ψ
[
�
]
�
�
[
�
]
.
\Psi [\gamma ]=\int [dA]\Psi [A]W_{\gamma }[A].
This defines the loop representation. Given an operator
�
^{\hat {O}} in the connection representation,
Φ
[
�
]
=
�
^
Ψ
[
�
]
�
�
1
,
{\displaystyle \Phi [A]={\hat {O}}\Psi [A]\qquad Eq\;1,}
one should define the corresponding operator
�
^
′
{\hat {O}}' on
Ψ
[
�
]
\Psi [\gamma ] in the loop representation via,
Φ
[
�
]
=
�
^
′
Ψ
[
�
]
�
�
2
,
{\displaystyle \Phi [\gamma ]={\hat {O}}'\Psi [\gamma ]\qquad Eq\;2,}
where
Φ
[
�
]
\Phi [\gamma ] is defined by the usual inverse loop transform,
Φ
[
�
]
=
∫
[
�
�
]
Φ
[
�
]
�
�
[
�
]
�
�
3.
\Phi [\gamma ]=\int [dA]\Phi [A]W_{\gamma }[A]\qquad Eq\;3.
A transformation formula giving the action of the operator
�
^
′
{\hat {O}}' on
Ψ
[
�
]
\Psi [\gamma ] in terms of the action of the operator
�
^{\hat {O}} on
Ψ
[
�
]
\Psi [A] is then obtained by equating the R.H.S. of
�
�
2
Eq\;2 with the R.H.S. of
�
�
3
Eq\;3 with
�
�
1
Eq\;1 substituted into
�
�
3
Eq\;3, namely
�
^
′
Ψ
[
�
]
=
∫
[
�
�
]
�
�
[
�
]
�
^
Ψ
[
�
]
,
{\hat {O}}'\Psi [\gamma ]=\int [dA]W_{\gamma }[A]{\hat {O}}\Psi [A],
or
�
^
′
Ψ
[
�
]
=
∫
[
�
�
]
(
�
^
†
�
�
[
�
]
)
Ψ
[
�
]
,
{\hat {O}}'\Psi [\gamma ]=\int [dA]({\hat {O}}^{\dagger }W_{\gamma }[A])\Psi [A],
where
�
^
†{\hat {O}}^{\dagger } means the operator
�
^{\hat {O}} but with the reverse factor ordering (remember from simple quantum
mechanics where the product of operators is reversed under conjugation). The action
of this operator on the Wilson loop is evaluated as a calculation in the connection
representation and the result is rearranged purely as a manipulation in terms of
loops (with regard to the action on the Wilson loop, the chosen transformed
operator is the one with the opposite factor ordering compared to the one used for
its action on wavefunctions
Ψ
[
�
]
\Psi [A]). This gives the physical meaning of the operator
�
^
′
{\hat {O}}'. For example, if
�
^
†{\displaystyle {\hat {O}}^{\dagger }} corresponded to a spatial diffeomorphism,
then this can be thought of as keeping the connection field
�
A of
�
�
[
�
]
W_{\gamma }[A] where it is while performing a spatial diffeomorphism on
�\gamma instead. Therefore, the meaning of
�
^
′
{\hat {O}}' is a spatial diffeomorphism on
�\gamma , the argument of
Ψ
[
�
]
\Psi [\gamma ].
�
~
^
†
�
�
[
�
]
=
−
�
�
�
�
�
^
�
�
�
�
�
�
�
�
�
�
�
�
�
�
�
[
�
]
.
{\displaystyle {\hat {\tilde {H}}}^{\dagger }W_{\gamma }[A]=-\epsilon _{ijk}{\hat
{F}}_{ab}^{k}{\frac {\delta }{\delta A_{a}^{i}}}{\frac {\delta }{\delta
A_{b}^{j}}}W_{\gamma }[A].}
When a derivative is taken it brings down the tangent vector,
�
˙
�
{\dot {\gamma }}^{a}, of the loop,
�\gamma . So,
�
^
�
�
�
�
˙
�
�
˙
�
.
{\displaystyle {\hat {F}}_{ab}^{i}{\dot {\gamma }}^{a}{\dot {\gamma }}^{b}.}
However, as
�
�
�
�
F_{ab}^{i} is anti-symmetric in the indices
�
a and
�
b this vanishes (this assumes that
�\gamma is not discontinuous anywhere and so the tangent vector is unique).
Geometric operators, the need for intersecting Wilson loops and spin network states
The easiest geometric quantity is the area. Let us choose coordinates so that the
surface
Σ\Sigma is characterized by
�
3
=
0
x^{3}=0. The area of small parallelogram of the surface
Σ\Sigma is the product of length of each side times
sin
�\sin \theta where
�\theta is the angle between the sides. Say one edge is given by the vector
�
→{\vec {u}} and the other by
�
→{\vec {v}} then,
�
=
‖
�
→
‖
‖
�
→
‖
sin
�
=
‖
�
→
‖
2
‖
�
→
‖
2
(
1
−
cos
2
�
)
=
‖
�
→
‖
2
‖
�
→
‖
2
−
(
�
→
⋅
�
→
)
2
{\displaystyle A=\|{\vec {u}}\|\|{\vec {v}}\|\sin \theta ={\sqrt {\|{\vec {u}}\|
^{2}\|{\vec {v}}\|^{2}(1-\cos ^{2}\theta )}}={\sqrt {\|{\vec {u}}\|^{2}\|{\vec
{v}}\|^{2}-({\vec {u}}\cdot {\vec {v}})^{2}}}}
In the space spanned by
�
1
x^{1} and
�
2
x^{2} there is an infinitesimal parallelogram described by
�
→
=
�
→
1
�
�
1
{\vec {u}}={\vec {e}}_{1}dx^{1} and
�
→
=
�
→
2
�
�
2
{\vec {v}}={\vec {e}}_{2}dx^{2}. Using
�
�
�
(
2
)
=
�
→
�
⋅
�
→
�
q_{AB}^{(2)}={\vec {e}}_{A}\cdot {\vec {e}}_{B} (where the indices
�
A and
�
B run from 1 to 2), yields the area of the surface
Σ\Sigma given by
�
Σ
=
∫
Σ
�
�
1
�
�
2
det
(
�
(
2
)
)
{\displaystyle A_{\Sigma }=\int _{\Sigma }dx^{1}dx^{2}{\sqrt {\det \left(q^{(2)}\
right)}}}
where
det
(
�
(
2
)
)
=
�
11
�
22
−
�
12
2
{\displaystyle \det(q^{(2)})=q_{11}q_{22}-q_{12}^{2}} and is the determinant of the
metric induced on
Σ\Sigma . The latter can be rewritten
det
(
�
(
2
)
)
=
�
�
�
�
�
�
�
�
�
�
�
�
/
2
{\displaystyle \det(q^{(2)})=\epsilon ^{AB}\epsilon ^{CD}q_{AC}q_{BD}/2} where the
indices
�
…
�
A\dots D go from 1 to 2. This can be further rewritten as
det
(
�
(
2
)
)
=
�
3
�
�
�
3
�
�
�
�
�
�
�
�
2
.
{\displaystyle \det(q^{(2)})={\epsilon ^{3ab}\epsilon ^{3cd}q_{ac}q_{bc} \over 2}.}
The standard formula for an inverse matrix is
�
�
�
=
�
�
�
�
�
�
�
�
�
�
�
�
�
�
2
!
det
(
�
)
.
{\displaystyle q^{ab}={\epsilon ^{bcd}\epsilon ^{aef}q_{ce}q_{df} \over 2!\
det(q)}.}
There is a similarity between this and the expression for
det
(
�
(
2
)
)
{\displaystyle \det(q^{(2)})}. But in Ashtekar variables,
�
~
�
�
�
~
�
�
=
det
(
�
)
�
�
�
{\displaystyle {\tilde {E}}_{i}^{a}{\tilde {E}}^{bi}=\det(q)q^{ab}}. Therefore,
�
Σ
=
∫
Σ
�
�
1
�
�
2
�
~
�
3
�
~
3
�
.
{\displaystyle A_{\Sigma }=\int _{\Sigma }dx^{1}dx^{2}{\sqrt {{\tilde {E}}_{i}^{3}
{\tilde {E}}^{3i}}}.}
According to the rules of canonical quantization the triads
�
~
�
3
{\tilde {E}}_{i}^{3} should be promoted to quantum operators,
�
~
^
�
3
∼
�
�
�
3
�
.
{\hat {{\tilde {E}}}}_{i}^{3}\sim {\delta \over \delta A_{3}^{i}}.
The area
�
ΣA_{\Sigma } can be promoted to a well defined quantum operator despite the fact
that it contains a product of two functional derivatives and a square-root.[13]
Putting
�
=
2
�
N=2J (
�
J-th representation),
∑
�
�
�
�
�
=
�
(
�
+
1
)
1.
{\displaystyle \sum _{i}T^{i}T^{i}=J(J+1)1.}
This quantity is important in the final formula for the area spectrum. The result
is
�
^
Σ
�
�
[
�
]
=
8
�
ℓ
Planck
2
�
∑
�
�
�
(
�
�
+
1
)
�
�
[
�
]
{\hat {A}}_{\Sigma }W_{\gamma }[A]=8\pi \ell _{\text{Planck}}^{2}\beta \sum _{I}{\
sqrt {j_{I}(j_{I}+1)}}W_{\gamma }[A]
where the sum is over all edges
�
I of the Wilson loop that pierce the surface
Σ\Sigma .
�
=
∫
�
�
3
�
det
(
�
)
=
∫
�
�
�
3
1
3
!
�
�
�
�
�
�
�
�
�
~
�
�
�
~
�
�
�
~
�
�
.
{\displaystyle V=\int _{R}d^{3}x{\sqrt {\det(q)}}=\int _{R}dx^{3}{\sqrt {{\frac {1}
{3!}}\epsilon _{abc}\epsilon ^{ijk}{\tilde {E}}_{i}^{a}{\tilde {E}}_{j}^{b}{\tilde
{E}}_{k}^{c}}}.}
The quantization of the volume proceeds the same way as with the area. Each time
the derivative is taken, it brings down the tangent vector
�
˙
�
{\dot {\gamma }}^{a}, and when the volume operator acts on non-intersecting Wilson
loops the result vanishes. Quantum states with non-zero volume must therefore
involve intersections. Given that the anti-symmetric summation is taken over in the
formula for the volume, it needs intersections with at least three non-coplanar
lines. At least four-valent vertices are needed for the volume operator to be non-
vanishing.
Tr
(
�
)
Tr
(
�
)
=
Tr
(
�
�
)
+
Tr
(
�
�
−
1
)
.
{\displaystyle \operatorname {Tr} (\mathbb {A} )\operatorname {Tr} (\mathbb {B} )=\
operatorname {Tr} (\mathbb {A} \mathbb {B} )+\operatorname {Tr} (\mathbb {A} \
mathbb {B} ^{-1}).}
This implies that given two loops
�\gamma and
�\eta that intersect,
�
�
[
�
]
�
�
[
�
]
=
�
�
∘
�
[
�
]
+
�
�
∘
�
−
1
[
�
]
W_{\gamma }[A]W_{\eta }[A]=W_{\gamma \circ \eta }[A]+W_{\gamma \circ \eta ^{-1}}[A]
where by
�
−
1
\eta ^{-1} we mean the loop
�\eta traversed in the opposite direction and
�
∘
�\gamma \circ \eta means the loop obtained by going around the loop
�\gamma and then along
�\eta . See figure below. Given that the matrices are unitary one has that
�
�
[
�
]
=
�
�
−
1
[
�
]
W_{\gamma }[A]=W_{\gamma ^{-1}}[A]. Also given the cyclic property of the matrix
traces (i.e.
Tr
(
�
�
)
=
Tr
(
�
�
)
{\displaystyle \operatorname {Tr} (\mathbb {A} \mathbb {B} )=\operatorname {Tr} (\
mathbb {B} \mathbb {A} )}) one has that
�
�
∘
�
[
�
]
=
�
�
∘
�
[
�
]
W_{\gamma \circ \eta }[A]=W_{\eta \circ \gamma }[A]. These identities can be
combined with each other into further identities of increasing complexity adding
more loops. These identities are the so-called Mandelstam identities. Spin networks
certain are linear combinations of intersecting Wilson loops designed to address
the over-completeness introduced by the Mandelstam identities (for trivalent
intersections they eliminate the over-completeness entirely) and actually
constitute a basis for all gauge invariant functions.
Finding the states that are annihilated by these constraints (the physical states),
and finding the corresponding physical inner product, and observables is the main
goal of the technical side of LQG.
Spin foams
Main articles: spin network, spin foam, BF model, and Barrett–Crane model
In loop quantum gravity (LQG), a spin network represents a "quantum state" of the
gravitational field on a 3-dimensional hypersurface. The set of all possible spin
networks (or, more accurately, "s-knots" – that is, equivalence classes of spin
networks under diffeomorphisms) is countable; it constitutes a basis of LQG Hilbert
space.
This then naturally gives rise to the two-complex (a combinatorial set of faces
that join along edges, which in turn join on vertices) underlying the spin foam
description; we evolve forward an initial spin network sweeping out a surface, the
action of the Hamiltonian constraint operator is to produce a new planar surface
starting at the vertex. We are able to use the action of the Hamiltonian constraint
on the vertex of a spin network state to associate an amplitude to each
"interaction" (in analogy to Feynman diagrams). See figure below. This opens up a
way of trying to directly link canonical LQG to a path integral description. Now
just as a spin networks describe quantum space, each configuration contributing to
these path integrals, or sums over history, describe 'quantum space-time'. Because
of their resemblance to soap foams and the way they are labeled John Baez gave
these 'quantum space-times' the name 'spin foams'.
The action of the Hamiltonian constraint translated to the path integral or so-
called spin foam description. A single node splits into three nodes, creating a
spin foam vertex.
�
(
�
�
)
N(x_{n}) is the value of
�
N at the vertex and
�
�
�
�
H_{nop} are the matrix elements of the Hamiltonian constraint
�
^{\hat {H}}.
There are however severe difficulties with this particular approach, for example
the Hamiltonian operator is not self-adjoint, in fact it is not even a normal
operator (i.e. the operator does not commute with its adjoint) and so the spectral
theorem cannot be used to define the exponential in general. The most serious
problem is that the
�
^
(
�
)
{\hat {H}}(x)'s are not mutually commuting, it can then be shown the formal
quantity
∫
[
�
�
]
�
�
∫
�
3
�
�
(
�
)
�
^
(
�
)
{\textstyle \int [dN]e^{i\int d^{3}xN(x){\hat {H}}(x)}} cannot even define a
(generalized) projector. The master constraint (see below) does not suffer from
these problems and as such offers a way of connecting the canonical theory to the
path integral formulation.
�
�
�
�
�
=
1
2
(
�
�
�
�
�
�
−
�
�
�
�
�
�
)
{\displaystyle B_{ab}^{IJ}={1 \over 2}\left(E_{a}^{I}E_{b}^{J}-E_{b}^{I}E_{a}^{J}\
right)}
(tetrads are like triads but in four spacetime dimensions), one recovers general
relativity. The condition that the
�
B field be given by the product of two tetrads is called the simplicity constraint.
The spin foam dynamics of the topological field theory is well understood. Given
the spin foam 'interaction' amplitudes for this simple theory, one then tries to
implement the simplicity conditions to obtain a path integral for general
relativity. The non-trivial task of constructing a spin foam model is then reduced
to the question of how this simplicity constraint should be imposed in the quantum
theory. The first attempt at this was the famous Barrett–Crane model.[16] However
this model was shown to be problematic, for example there did not seem to be enough
degrees of freedom to ensure the correct classical limit.[17] It has been argued
that the simplicity constraint was imposed too strongly at the quantum level and
should only be imposed in the sense of expectation values just as with the Lorenz
gauge condition
∂
�
�
^
�\partial _{\mu }{\hat {A}}^{\mu } in the Gupta–Bleuler formalism of quantum
electrodynamics. New models have now been put forward, sometimes motivated by
imposing the simplicity conditions in a weaker sense.
Much progress has been made with regard to this issue by Engle, Pereira, and
Rovelli,[18] Freidel and Krasnov[19] and Livine and Speziale[20] in defining spin
foam interaction amplitudes with much better behaviour.
An attempt to make contact between EPRL-FK spin foam and the canonical formulation
of LQG has been made.[21]
Concerning issue number 2 above, one can consider so-called weave states. Ordinary
measurements of geometric quantities are macroscopic, and planckian discreteness is
smoothed out. The fabric of a T-shirt is analogous: at a distance it is a smooth
curved two-dimensional surface, but on closer inspection we see that it is actually
composed of thousands of one-dimensional linked threads. The image of space given
in LQG is similar. Consider a very large spin network formed by a very large number
of nodes and links, each of Planck scale. Probed at a macroscopic scale, it appears
as a three-dimensional continuous metric geometry.
To make contact with familiar low energy physics it is mandatory to have to develop
approximation schemes both for the physical inner product and for Dirac
observables; the spin foam models that have been intensively studied can be viewed
as avenues toward approximation schemes for said physical inner product.
�
(
�
1
,
…
,
�
�
)
=
⟨
0
|
�
(
�
�
)
…
�
(
�
1
)
|
0
⟩
,
{\displaystyle W(x_{1},\dots ,x_{n})=\langle 0|\phi (x_{n})\dots \phi (x_{1})|0\
rangle ,}
completely determine the theory. In particular, one can calculate the scattering
amplitudes from these quantities. As explained below in the section on the
Background independent scattering amplitudes, in the background-independent
context, the
�
−n- point functions refer to a state and in gravity that state can naturally encode
information about a specific geometry which can then appear in the expressions of
these quantities. To leading order, LQG calculations have been shown to agree in an
appropriate sense with the
�
−n-point functions calculated in the effective low energy quantum general
relativity.
That the master constraint Poisson algebra is an honest Lie algebra opens up the
possibility of using a certain method, known as group averaging, in order to
construct solutions of the infinite number of Hamiltonian constraints, a physical
inner product thereon and Dirac observables via what is known as refined algebraic
quantization, or RAQ.[30]
�
^
:=
∫
�
3
�
(
�
det
(
�
(
�
)
)
4
)
^
†
(
�
)
(
�
det
(
�
(
�
)
)
4
)
^
(
�
)
.
{\displaystyle {\hat {M}}:=\int d^{3}x{\widehat {\left({\frac {H}{\sqrt[{4}]{\
det(q(x))}}}\right)}}^{\dagger }(x){\widehat {\left({\frac {H}{\sqrt[{4}]{\
det(q(x))}}}\right)}}(x).}
Obviously,
(
�
det
(
�
(
�
)
)
4
)
^
(
�
)
Ψ
=
0
{\displaystyle {\widehat {\left({\frac {H}{\sqrt[{4}]{\det(q(x))}}}\right)}}(x)\Psi
=0}
for all
�
x implies
�
^
Ψ
=
0
{\hat {M}}\Psi =0. Conversely, if
�
^
Ψ
=
0
{\hat {M}}\Psi =0 then
0
=
⟨
Ψ
,
�
^
Ψ
⟩
=
∫
�
3
�
‖
(
�
det
(
�
(
�
)
)
4
)
^
(
�
)
Ψ
‖
2
�
�
4
{\displaystyle 0=\left\langle \Psi ,{\hat {M}}\Psi \right\rangle =\int d^{3}x\
left\|{\widehat {\left({\frac {H}{\sqrt[{4}]{\det(q(x))}}}\right)}}(x)\Psi \right\|
^{2}\qquad Eq\;4}
implies
(
�
det
(
�
(
�
)
)
4
)
^
(
�
)
Ψ
=
0
{\displaystyle {\widehat {\left({\frac {H}{\sqrt[{4}]{\det(q(x))}}}\right)}}(x)\Psi
=0}.
What is done first is, we are able to compute the matrix elements of the would-be
operator
�
^{\hat {M}}, that is, we compute the quadratic form
�
�
Q_{M}. It turns out that as
�
�
Q_{M} is a graph changing, diffeomorphism invariant quadratic form it cannot exist
on the kinematic Hilbert space
�
�
�
�
H_{{Kin}}, and must be defined on
�
�
�
�
�
H_{Diff}. Since the master constraint operator
�
^{\hat {M}} is densely defined on
�
�
�
�
�
H_{{Diff}}, then
�
^{\hat {M}} is a positive and symmetric operator in
�
�
�
�
�
H_{{Diff}}. Therefore, the quadratic form
�
�
Q_{M} associated with
�
^{\hat {M}} is closable. The closure of
�
�
Q_{M} is the quadratic form of a unique self-adjoint operator
�
¯
^{\hat {\overline {M}}}, called the Friedrichs extension of
�
^{\hat {M}}. We relabel
�
¯
^{\hat {\overline {M}}} as
�
^{\hat {M}} for simplicity.
Note that the presence of an inner product, viz Eq 4, means there are no
superfluous solutions i.e. there are no
Ψ\Psi such that
(
�
det
(
�
(
�
)
)
4
)
^
(
�
)
Ψ
≠
0
,
{\displaystyle {\widehat {\left({\frac {H}{\sqrt[{4}]{\det(q(x))}}}\right)}}(x)\Psi
\not =0,}
but for which
�
^
Ψ
=
0
{\hat {M}}\Psi =0.
The spectrum of the master constraint may not contain zero due to normal or factor
ordering effects which are finite but similar in nature to the infinite vacuum
energies of background-dependent quantum field theories. In this case it turns out
to be physically correct to replace
�
^{\hat {M}} with
�
^
′
:=
�
^
−
min
(
�
�
�
�
(
�
^
)
)
1
^{\displaystyle {\hat {M}}':={\hat {M}}-\min(spec({\hat {M}})){\hat {1}}} provided
that the "normal ordering constant" vanishes in the classical limit, that is,
lim
ℏ
→
0
min
(
�
�
�
�
(
�
^
)
)
=
0
,
{\displaystyle \lim _{\hbar \to 0}\min(spec({\hat {M}}))=0,}
so that
�
^
′
{\hat {M}}' is a valid quantisation of
�
M.
�
�
=
∫
Σ
�
3
�
�
(
�
)
2
−
�
�
�
�
�
(
�
)
�
�
(
�
)
det
(
�
)
{\displaystyle M_{E}=\int _{\Sigma }d^{3}x{H(x)^{2}-q^{ab}V_{a}(x)V_{b}(x) \over {\
sqrt {\det(q)}}}}.
Setting this single constraint to zero is equivalent to
�
(
�
)
=
0
H(x)=0 and
�
�
(
�
)
=
0
V_{a}(x)=0 for all
�
x in
Σ\Sigma . This constraint implements the spatial diffeomorphism and Hamiltonian
constraint at the same time on the Kinematic Hilbert space. The physical inner
product is then defined as
⟨
�
,
�
⟩
Phys
=
lim
�
→
∞
⟨
�
,
∫
−
�
�
�
�
�
�
�
�
^
�
�
⟩
{\displaystyle \langle \phi ,\psi \rangle _{\text{Phys}}=\lim _{T\to \infty }\left\
langle \phi ,\int _{-T}^{T}dte^{it{\hat {M}}_{E}}\psi \right\rangle }
(as
�
(
�
�
^
)
=
lim
�
→
∞
∫
−
�
�
�
�
�
�
�
�
^
�
{\textstyle \delta ({\hat {M_{E}}})=\lim _{T\to \infty }\int _{-T}^{T}dte^{it{\hat
{M}}_{E}}}). A spin foam representation of this expression is obtained by splitting
the
�
t-parameter in discrete steps and writing
�
�
�
�
^
�
=
lim
�
→
∞
[
�
�
�
�
^
�
/
�
]
�
=
lim
�
→
∞
[
1
+
�
�
�
^
�
/
�
]
�
.
{\textstyle e^{it{\hat {M}}_{E}}=\lim _{n\to \infty }\left[e^{it{\hat {M}}_{E}/n}\
right]^{n}=\lim _{n\to \infty }[1+it{\hat {M}}_{E}/n]^{n}.}
The spin foam description then follows from the application of
[
1
+
�
�
�
^
�
/
�
]
[1+it{\hat {M}}_{E}/n] on a spin network resulting in a linear combination of new
spin networks whose graph and labels have been modified. Obviously an approximation
is made by truncating the value of
�
n to some finite integer. An advantage of the extended master constraint is that we
are working at the kinematic level and so far it is only here we have access
semiclassical coherent states. Moreover, one can find none graph changing versions
of this master constraint operator, which are the only type of operators
appropriate for these coherent states.
An artist depiction of two black holes merging, a process in which the laws of
thermodynamics are upheld.
Black hole thermodynamics is the area of study that seeks to reconcile the laws of
thermodynamics with the existence of black hole event horizons. The no hair
conjecture of general relativity states that a black hole is characterized only by
its mass, its charge, and its angular momentum; hence, it has no entropy. It
appears, then, that one can violate the second law of thermodynamics by dropping an
object with nonzero entropy into a black hole.[46] Work by Stephen Hawking and
Jacob Bekenstein showed that one can preserve the second law of thermodynamics by
assigning to each black hole a black-hole entropy
�
BH
=
�
B
�
4
ℓ
P
2
,
S_{\text{BH}}={\frac {k_{\text{B}}A}{4\ell _{\text{P}}^{2}}},
where
�
A is the area of the hole's event horizon,
�
B
k_{\text{B}} is the Boltzmann constant, and
ℓ
P
=
�
ℏ
/
�
3
{\textstyle \ell _{\text{P}}={\sqrt {G\hbar /c^{3}}}} is the Planck length.[47] The
fact that the black hole entropy is also the maximal entropy that can be obtained
by the Bekenstein bound (wherein the Bekenstein bound becomes an equality) was the
main observation that led to the holographic principle.[46]
An oversight in the application of the no-hair theorem is the assumption that the
relevant degrees of freedom accounting for the entropy of the black hole must be
classical in nature; what if they were purely quantum mechanical instead and had
non-zero entropy? Actually, this is what is realized in the LQG derivation of black
hole entropy, and can be seen as a consequence of its background-independence – the
classical black hole spacetime comes about from the semiclassical limit of the
quantum state of the gravitational field, but there are many quantum states that
have the same semiclassical limit. Specifically, in LQG[48] it is possible to
associate a quantum geometrical interpretation to the microstates: These are the
quantum geometries of the horizon which are consistent with the area,
�
A, of the black hole and the topology of the horizon (i.e. spherical). LQG offers a
geometric explanation of the finiteness of the entropy and of the proportionality
of the area of the horizon.[49][50] These calculations have been generalized to
rotating black holes.[51]
A recent success of the theory in this direction is the computation of the entropy
of all non singular black holes directly from theory and independent of Immirzi
parameter.[52][53] The result is the expected formula
�
=
�
/
4
S=A/4, where
�
S is the entropy and
�
A the area of the black hole, derived by Bekenstein and Hawking on heuristic
grounds. This is the only known derivation of this formula from a fundamental
theory, for the case of generic non singular black holes. Older attempts at this
calculation had difficulties. The problem was that although Loop quantum gravity
predicted that the entropy of a black hole is proportional to the area of the event
horizon, the result depended on a crucial free parameter in the theory, the above-
mentioned Immirzi parameter. However, there is no known computation of the Immirzi
parameter, so it had to be fixed by demanding agreement with Bekenstein and
Hawking's calculation of the black hole entropy.
Based on the fluctuations of the horizon area, a quantum black hole exhibits
deviations from the Hawking spectrum that would be observable were X-rays from
Hawking radiation of evaporating primordial black holes to be observed.[54] The
quantum effects are centered at a set of discrete and unblended frequencies highly
pronounced on top of Hawking radiation spectrum.[55]
Planck star
Main articles: Planck star, black hole firewall, and black hole information paradox
In 2014 Carlo Rovelli and Francesca Vidotto proposed that there is a Planck star
inside every black hole.[56] Based on LQG, the theory states that as stars are
collapsing into black holes, the energy density reaches the Planck energy density,
causing a repulsive force that creates a star. Furthermore, the existence of such a
star would resolve the black hole firewall and black hole information paradox.
Achievements of LQC have been the resolution of the big bang singularity, the
prediction of a Big Bounce, and a natural mechanism for inflation.
LQC models share features of LQG and so is a useful toy model. However, the results
obtained are subject to the usual restriction that a truncated classical theory,
then quantized, might not display the true behaviour of the full theory due to
artificial suppression of degrees of freedom that might have large quantum
fluctuations in the full theory. It has been argued that singularity avoidance in
LQC are by mechanisms only available in these restrictive models and that
singularity avoidance in the full theory can still be obtained but by a more subtle
feature of LQG.[58][59]
Loop quantum gravity phenomenology
Quantum gravity effects are notoriously difficult to measure because the Planck
length is so incredibly small. However recently physicists, such as Jack Palmer,
have started to consider the possibility of measuring quantum gravity effects
mostly from astrophysical observations and gravitational wave detectors. The energy
of those fluctuations at scales this small cause space-perturbations which are
visible at higher scales.
A strategy for addressing this problem has been suggested;[60] the idea is to study
the boundary amplitude, namely a path integral over a finite space-time region,
seen as a function of the boundary value of the field.[61][62] In conventional
quantum field theory, this boundary amplitude is well–defined[63][64] and codes the
physical information of the theory; it does so in quantum gravity as well, but in a
fully background–independent manner.[65] A generally covariant definition of
�
n-point functions can then be based on the idea that the distance between physical
points –arguments of the
�
n-point function is determined by the state of the gravitational field on the
boundary of the spacetime region considered.
LQG never introduces a background and excitations living on this background, so LQG
does not use gravitons as building blocks. Instead one expects that one may recover
a kind of semiclassical limit or weak field limit where something like "gravitons"
will show up again. In contrast, gravitons play a key role in string theory where
they are among the first (massless) level of excitations of a superstring.
LQG differs from string theory in that it is formulated in 3 and 4 dimensions and
without supersymmetry or Kaluza–Klein extra dimensions, while the latter requires
both to be true. There is no experimental evidence to date that confirms string
theory's predictions of supersymmetry and Kaluza–Klein extra dimensions. In a 2003
paper "A Dialog on Quantum Gravity",[66] Carlo Rovelli regards the fact LQG is
formulated in 4 dimensions and without supersymmetry as a strength of the theory as
it represents the most parsimonious explanation, consistent with current
experimental results, over its rival string/M-theory. Proponents of string theory
will often point to the fact that, among other things, it demonstrably reproduces
the established theories of general relativity and quantum field theory in the
appropriate limits, which loop quantum gravity has struggled to do. In that sense
string theory's connection to established physics may be considered more reliable
and less speculative, at the mathematical level. Loop quantum gravity has nothing
to say about the matter (fermions) in the universe.
Since LQG has been formulated in 4 dimensions (with and without supersymmetry), and
M-theory requires supersymmetry and 11 dimensions, a direct comparison between the
two has not been possible. It is possible to extend mainstream LQG formalism to
higher-dimensional supergravity, general relativity with supersymmetry and Kaluza–
Klein extra dimensions should experimental evidence establish their existence. It
would therefore be desirable to have higher-dimensional Supergravity loop
quantizations at one's disposal in order to compare these approaches. In fact a
series of papers have been published attempting just this.[67][68][69][70][71][72]
[73][74] Most recently, Thiemann (and alumni) have made progress toward calculating
black hole entropy for supergravity in higher dimensions. It will be interesting to
compare these results to the corresponding super string calculations.[75][76]