You are on page 1of 15

uantum Gravity

First published Mon 26 Dec, 2005

Quantum Gravity: A physical theory describing the gravitational interactions of matter


and energy in which matter and energy are described by quantum theory. In most, but not
all, theories of quantum gravity, gravity is also quantized. Since the contemporary theory
of gravity, general relativity, describes gravitation as the curvature of spacetime by matter
and energy, a quantization of gravity implies some sort of quantization of spacetime
itself. Insofar as all extant physical theories rely on a classical spacetime background, this
presents profound methodological and ontological challenges for the philosopher and the
physicist.

• 1. Introduction
• 2. Gravity Meets Quantum Theory
• 3. Methodology
o 3.1 Theory
o 3.2 Experiment
• 4. Theoretical Frameworks
o 4.1 String theory
o 4.2 Canonical and loop quantum gravity
 4.2.1 Geometric variables
 4.2.2 Problem of time
 4.2.3 Ashtekar, loop, and other variables
o 4.3 Other approaches
• 5. Philosophical Issues
o 5.1 Time
o 5.2 Ontology
o 5.3 Status of quantum theory
o 5.4 Methodology
• 6. Conclusion
• Bibliography
• Other Internet Resources
• Related Entries

1. Introduction
Dutch artist M.C. Escher's elegant pictorial paradoxes are prized by many, not least by
philosophers, physicists, and mathematicians. Some of his work, for example Ascending
and Descending, relies on optical illusion to depict what is actually an impossible
situation. Other works are paradoxical in the broad sense, but not impossible: Relativity
depicts a coherent arrangement of objects, albeit an arrangement in which the force of
gravity operates in an unfamiliar fashion. (See the Other Internet Resources section
below for images.) Quantum gravity itself may be like this: an unfamiliar yet coherent
arrangement of familiar elements. Or it may be like Ascending and Descending, an
impossible construction which looks sensible in its local details but does not fit together
into a coherent whole.

‘Quantum gravity’ primarily refers to an area of research, rather than a particular theory
of quantum gravity. Several approaches exist, none of them entirely successful to date.
Thus the philosopher's task, if indeed she has one, is different from what it is when
dealing with a more-or-less settled body of theory such as classical Newtonian
mechanics, general relativity, or quantum mechanics. In such cases, one typically
proceeds by assuming the validity of the theory or theoretical framework and drawing the
ontological and perhaps epistemological consequences of the theory, trying to understand
what it is that the theory is telling us about the nature of space, time, matter, causation,
and so on. Theories of quantum gravity, on the other hand, are bedeviled by a host of
technical and conceptual problems, questions, and issues which make them unsuited to
this approach. However, philosophers who have a taste for a broader and more open-
ended form of inquiry will find much to think about.

2. Gravity Meets Quantum Theory


The difficulties in reconciling quantum theory and gravity into some form of quantum
gravity come from the prima facie incompatibility of general relativity, Einstein's
relativistic theory of gravitation, with quantum field theory, the framework for the
description of the other three forces (electromagnetism and the strong and weak nuclear
interactions). Whence the incompatibility? General relativity is described by Einstein's
equations, which amount to constraints on the curvature of spacetime (the Einstein tensor
on the left-hand side) due to the presence of mass and other forms of energy, such as
electromagnetic radiation (the stress-energy-momentum tensor on the right-hand side).
(See John Baez's webpages in Other Internet Resources for an excellent introduction.) In
doing so, they manage to encompass traditional, Newtonian gravitational phenomena
such as the mutual attraction of two or more massive objects, while also predicting new
phenomena such as the bending of light by these objects (which has been observed) and
the existence of gravitational radiation (which has to date only been indirectly observed
via the decrease in the period of binary pulsars). (For the latter observation, see the 1993
Physics Nobel Prize presentation speech by Carl Nordling .)

In general relativity, mass and energy are treated in a purely classical manner, where
‘classical’ means that physical quantities such as the strengths and directions of various
fields and the positions and velocities of particles have definite values. These quantities
are represented by tensor fields, sets of (real) numbers associated with each spacetime
point. For example, the stress, energy, and momentum Tab(x,t) of the electromagnetic
field at some point (x,t), are functions of the three components Ei, Ej, Ek, Bi, Bj, Bk of the
electric and magnetic fields E and B at that point. These quantities in turn determine, via
Einstein's equations, an aspect of the ‘curvature’ of spacetime, a set of numbers Gab(x,t)
which is in turn a function of the spacetime metric. The metric gab(x,t) is a set of numbers
associated with each point which gives the distance to neighboring points. At the end of
the day, a model of the world according to general relativity consists of a spacetime
manifold with a metric, the curvature of which is constrained by the stress-energy-
momentum of the matter distribution. All physical quantities — the value of the x-
component of the electric field at some point, the scalar curvature of spacetime at some
point — have definite values, given by real (as opposed to complex or imaginary)
numbers. Thus general relativity is a classical theory in the sense given above.

The problem is that our fundamental theories of matter and energy, the theories
describing the interactions of various particles via the electromagnetic force and the
strong and weak nuclear forces, are all quantum theories. In quantum theories, these
physical quantities do not in general have definite values. For example, in quantum
mechanics, the position of an electron may be specified with arbitrarily high accuracy
only at the cost of a loss of specificity in the description of its momentum, hence its
velocity. At the same time, in the quantum theory of the electromagnetic field known as
quantum electrodynamics (QED), the electric and magnetic fields associated with the
electron suffer an associated uncertainty. In general, physical quantities are described by
a quantum state which gives a probability distribution over many different values, and
increased specificity (narrowing of the distribution) of one property (e.g., position,
electric field) gives rise to decreased specificity of its canonically conjugate property
(e.g., momentum, magnetic field). This is an expression of Heisenberg's Uncertainty
Principle.

On the surface, the incompatibility between general relativity and quantum theory might
seem rather trivial. Why not just follow the model of QED and quantize the gravitational
field, similar to the way in which the electromagnetic field was quantized? Just as we
associate a quantum state of the electromagnetic field with the quantum state of
electrically charged matter, we should, one might think, similarly just associate a
quantum state of the gravitational field with the quantum state of both charged and
uncharged matter. This is more or less the path that has been taken, but it encounters
extraordinary difficulties. Some physicists consider these to be technical difficulties,
having to do with the non-renormalizability of the gravitational interaction and the
consequent failure of the perturbative methods which have proven effective in ordinary
quantum field theories. However, these technical problems are closely related to a set of
daunting conceptual difficulties, of interest to both physicists and philosophers.

The conceptual difficulties basically follow from the nature of the gravitational
interaction, in particular the equivalence of gravitational and inertial mass, which allows
one to represent gravity as a property of spacetime itself, rather than as a field
propagating in a (passive) spacetime background. When one attempts to quantize gravity,
one is subjecting some of the properties of spacetime to quantum fluctuations. But
ordinary quantum theory presupposes a well-defined classical background against which
to define these fluctuations (Weinstein, 2001a, b), and so one runs into trouble not only in
giving a mathematical characterization of the quantization procedure (how to take into
account these fluctuations in the effective spacetime structure?) but also in giving a
conceptual and physical account of the theory that results, should one succeed. We will
look in more detail at how these conceptual problems arise in two different research
programs below. But first, we will talk a bit about some general methodological issues
which haunt the field.

3. Methodology
Research in quantum gravity has a rather peculiar flavor, owing to both the technical and
conceptual difficulty of the field and the remoteness from experiment. Thus conventional
notions of the relation between theory and experiment have a tenuous foothold, at best.

3.1 Theory

As remarked in the introduction, there is no single, generally agreed-upon body of theory


in quantum gravity. The majority of the physicists working in the field work on string
theory, an ambitious program which aims at providing a unified theory of all interactions.
A non-negligible minority work on what is now called loop quantum gravity, the goal of
which is simply to provide a quantum theory of the gravitational interaction. There is also
significant work in other areas. [Good recent reviews of the theoretical landscape include
Carlip 2001 and Smolin 2001 (Other Internet Resources section below), 2003.] But there
is no real consensus, for two reasons.

The first reason is that it is extremely difficult to make any concrete predictions in these
theories. String theory, in particular, is plagued by a lack of predictions because of the
tremendous number of distinct ground or vacuum states in the theory, with an absence of
guiding principles for singling out the physically significant ones. Though the string
community prides itself on the dearth of free parameters in the theory (in contrast to the
nineteen or so free parameters found in the standard model of particle physics), the
problem arguably resurfaces in the huge number of vacua associated with different
compactifications of the nine space dimensions to the three we observe. Attempts to
explain why we live in the particular vacuum that we do have recently given rise to
appeals to the infamous anthropic principle (Susskind, 2003 [CHECK THE DATE]),
whereby the existence of humans is invoked to, in some sense, “explain” the fact that we
find ourselves in a particular world.

Loop quantum gravity is less plagued by a lack of predictions, and indeed it is often
claimed that the discreteness of area and volume are concrete predictions of the theory.
Proponents of this approach argue that this makes the theory more susceptible to
falsification, thus more scientific (in the sense of Popper; see the entry on Karl Popper)
than string theory. However, it is still quite unclear, in practice and even in principle, how
one might actually observe these quantities.

3.2 Experiment

The second reason for the absence of consensus is that there are no experiments in
quantum gravity, and little in the way of observations that might qualify as direct or
indirect data or evidence. This stems in part from the lack of theoretical predictions, since
it is difficult to design an observational test of a theory if one does not know where to
look or what to look at. But it also stems from the fact that most theories of quantum
gravity appear to predict departures from classical relativity only at energy scales on the
order of 1019 GeV. (By way of comparison, the proton-proton collisions at Fermilab have
an energy on the order of 103 GeV.) Whereas research in particle physics proceeds in
large part by examining the data collected in large particle accelerators, accelerators
which are able to smash particles together at sufficiently high energies to probe the
properties of atomic nuclei, gravity is so weak that there is no way to do a comparable
experiment that would reveal properties at the energy scales at which quantum
gravitational effects are expected to be important.

Though progress is being made in trying to at least draw observational consequences of


loop quantum gravity, a theory of quantum gravity which arguably does make predictions
(Amelino-Camelia, 2003, in the Other Internet Resources section below; D. Mattingly,
2005), it is remarkable that the most notable “test” of quantum theories of gravity
imposed by the community to date involves a phenomenon which has never been
observed, the so-called Hawking radiation from black holes. Based on earlier work of
Bekenstein (1973) and others, Hawking (1974) predicted that black holes would radiate
energy, and would do so in proportion to their gravitational “temperature,” which was in
turn understood to be proportional to their mass, angular momentum, and charge.
Associated with this temperature is an entropy (see the entry on the philosophy of
statistical mechanics), and one would expect a theory of quantum gravity to allow one to
calculate the entropy associated with a black hole of given mass, angular momentum, and
charge, the entropy corresponding to the number of quantum states of the gravitational
field having the same mass, charge, and angular momentum. (See Unruh (2001) and
references therein.)

In their own ways, string theory and loop quantum gravity have both passed the test of
predicting an entropy for black holes which accords with Hawking's calculation. String
theory gets the number right for a not-particularly physically realistic subset of black
holes called near-extremal black holes, while loop quantum gravity gets it right for
generic black holes, but only up to an overall constant. If the Hawking effect is real, then
this consonance could be counted as evidence in favor of either or both theories.

It should be noted, finally, that to date neither of the main research programs has been
shown to give rise to the world we see at low energies. Indeed, it is a major challenge of
loop quantum gravity to show that it has general relativity as a low-energy limit, and a
major challenge of string theory to show that it has the standard model of particle physics
plus general relativity as a low-energy limit.

The absence of relevant experiments in quantum gravity is a peculiarity which has drawn
little attention to date from philosophers of science (an exception is Butterfield & Isham,
2001). However, it would seem to be fertile ground for philosophical analysis, in that it
raises the interesting question of how a science should and does proceed in the absence of
data.

4. Theoretical Frameworks
4.1 String Theory

Known variously as string theory, superstring theory, and M-theory, this program has its
roots, indirectly, in the observation, dating back to at least the 1950s, that classical
general relativity looks in many ways like the theory of a massless ‘spin-two’ field
propagating on the flat Minkowski spacetime of special relativity. [See Rovelli 2001b
(Other Internet Resources section below), and 2006 for a capsule history, and Greene
2000 for a popular account.] This observation led to early attempts to formulate a
quantum theory of gravity by “quantizing” this spin-two theory. However, it turned out
that the theory is not perturbatively renormalizable, meaning that there are ineliminable
infinities. Attempts to modify the classical theory to eliminate this problem led to a
different problem, non-unitarity, and so this general approach was moribund until the
mid-1970s, when it was discovered that a theory of one-dimensional “strings” developed
around 1970 to account for the strong interaction, actually provided a framework for a
unified theory which included gravity, because one of the modes of oscillation of the
string corresponded to a massless spin-two particle (the ‘graviton’).

The original and still prominent idea behind string theory was to replace the point
particles of ordinary quantum field theory (particles like photons, electrons, etc) with
one-dimensional extended objects called strings. (See Weingard, 2001 and Witten, 2001
for overviews of the conceptual framework.) In the early development of the theory, it
was recognized that construction of a consistent quantum theory of strings required that
the strings “live” in a larger number of spatial dimensions than the observed three;
eventually, most string theories came to be formulated in nine space dimensions and one
time dimension. Strings can be open or closed, and have a characteristic tension and
hence vibrational spectrum. The various modes of vibration correspond to various
particles, one of which is the graviton. The resulting theories have the advantage of being
perturbatively renormalizable, at least to second order. This means that perturbative
calculations are at least mathematically tractable. Since perturbation theory is an almost
indispensable tool for physicists, this is deemed a good thing.

String theory has undergone several mini-revolutions over the last several years, one of
which involved the discovery of various duality relations, mathematical transformations
connecting, in this case, what appeared to be mathematically distinct string theories —
type I, type IIA, type IIB, HE and HO — to one another and to eleven-dimensional
supergravity (a particle theory). The discovery of these connections led to the conjecture
that all of the string theories are really aspects of a single underlying theory, which was
given the name ‘M-theory’ (though M-theory is also used more specifically to describe
the unknown theory of which eleven-dimensional supergravity is the low energy limit).
The rationale is that what looks like one theory at strong coupling (high energy
description) looks like another theory at weak coupling (lower energy, more tractable
description), and that if all the theories are related to one another, then they must all be
aspects of some more fundamental theory. Though attempts have been made, there has
been no successful formulation of this theory: its very existence, much less its nature, is
still largely a matter of conjecture.
4.2 Canonical and Loop Quantum Gravity

Whereas string theory views the curved spacetime of general relativity as an effective
modification of a flat (or other fixed) background geometry by a massless spin-two field,
the canonical quantum gravity program treats the spacetime metric itself as a kind of
field, and attempts to quantize it directly.

Technically, most work in this camp proceeds by writing down general relativity in so-
called ‘canonical’ or ‘Hamiltonian’ form, since there is a more-or-less clearcut way to
quantize theories once they are put in this form (Kuchar, 1993; Belot & Earman, 2001).
In a canonical description, one chooses a particular set of configuration variables xi and
canonically conjugate momentum variables pi which describe the state of a system at
some time. Then, one obtains the time-evolution of these variables from the Hamiltonian
H(xi,pi). Quantization proceeds by treating the configuration and momentum variables as
operators on a quantum state space (a Hilbert space) obeying certain commutation
relations analogous to the classical Poisson-bracket relations, which effectively encode
the quantum fuzziness associated with Heisenberg's uncertainty principle.

Although advocates of the canonical approach often accuse string theorists of relying too
heavily on classical background spacetime, the canonical approach does something which
is arguably quite similar, in that one begins with a theory that conceives time-evolution in
terms of evolving some data given on a spacelike surface, and then quantizing the theory.
The problem is that if spacetime is quantized, this assumption does not make sense in
anything but an approximate way. This issue in particular is decidedly neglected in both
the physical and philosophical literature (but see Isham (1993)), and there is more that
might be said.

4.2.1 Geometric variables

Early attempts at quantizing general relativity by Dirac, Wheeler, DeWitt and others in
the 1950s and 1960s worked with a seemingly natural choice for configuration variables,
namely geometric variables gij corresponding to the various components of the ‘three-
metric’ describing the intrinsic geometry of the given spatial slice of spacetime. One can
think about arriving at this via an arbitrary slicing of a 4-dimensional “block” universe by
3-dimensional spacelike hypersurfaces. The conjugate momenta πij then effectively
encode the time rate-of-change of the metric, which, from the 4-dimensional perspective,
is directly related to the extrinsic curvature of the slice (meaning the curvature relative to
the spacetime in which the slice is embedded). This approach is known as
‘geometrodynamics’.

In these geometric variables, as in any other canonical formulation of general relativity,


one is faced with constraints, which encode the fact that the canonical variables cannot be
specified independently. A familiar example of a constraint is Gauss's law from ordinary
electromagnetism, which states that, in the absence of charges, ∇·E(x) = 0 at every point
x. It means that the three components of the electric field at every point must be chosen
so as to satisfy this constraint, which in turn means that there are only two “true” degrees
of freedom possessed by the electric field at any given point in space. (Specifying two
components of the electric field at every point dictates the third component.)

The constraints in electromagnetism may be viewed as stemming from the U(1) gauge
invariance of Maxwell's theory, while the constraints of general relativity stem from the
diffeomorphism invariance of the theory. Diffeomorphism invariance means, informally,
that one can take a solution of Einstein's equations and drag it (meaning the metric and
the matter fields) around on the spacetime manifold and obtain a mathematically distinct
but physically equivalent solution. The three ‘supermomentum’ constraints in the
canonical theory reflect the freedom to drag the metric and matter fields around in
various directions on a given three-dimensional spacelike hypersurface, while the ‘super-
Hamiltonian’ constraint reflects the freedom to drag the fields in the “time” direction, and
so to the “next” hypersurface. (Each constraint applies at each point of the given
spacelike hypersurface, so that there are actually 4 × ∞3 constraints, four for each point.)
In the classical (unquantized) canonical formulation of general relativity, the constraints
do not pose any particular conceptual problems. One effectively chooses a background
space and time (via a choice of the lapse and shift functions) “on the fly”, and one can be
confident that the spacetime that results is independent of the particular choice.
Effectively, different choices of these functions give rise to different choices of
background against which to evolve the foreground. However, the constraints pose a
serious problem when one moves to quantum theory.

4.2.2 Problem of time

All approaches to canonical quantum gravity face the so-called “problem of time” in one
form or another (Kuchař (1992) and Isham (1993) are excellent reviews). The problem
stems from the fact that in preserving the diffeomorphism-invariance of general relativity
— depriving the coordinates of the background manifold of any physical meaning — the
“slices” of spacetime one is considering inevitably include time, just as they include
space. In the canonical formulation, the diffeomorphism invariance is reflected in the
constraints, and the inclusion of what would ordinarily be a ‘time’ variable in the data is
reflected in the existence of the super-Hamiltonian constraint. The difficulties presented
by this constraint constitute the problem of time.

Attempts to quantize general relativity in the canonical framework proceed by turning the
canonical variables into operators on an appropriate state space (e.g., the space of square-
integrable functions over three-metrics), and dealing somehow with the constraints.
When quantizing a theory with constraints, there are two possible approaches. The
approach usually adopted in gauge theories is to deal with the constraints before
quantization, so that only true degrees of freedom are promoted to operators when
passing to the quantum theory. There are a variety of ways of doing this so-called ‘gauge
fixing’, but they all involve removing the extra degrees of freedom by imposing some
special conditions. In general relativity, fixing a gauge is tantamount to specifying a
particular coordinate system with respect to which the “physical” data is described
(spatial coordinates) and with respect to which it evolves (time coordinate). This is
difficult already at the classical level, since the utility and, moreover, the very tractability
of any particular gauge generally depends on the properties of the solution to the
equations, which of course is what one is trying to find in the first place. But in the
quantum theory, one is faced with the additional concern that the resulting theory may
well not be independent of the choice of gauge. This is closely related to the problem of
identifying true, gauge-invariant observables in the classical theory (Torre 2005, in the
Other Internet Resources section).

The preferred approach in canonical quantum gravity is to impose the constraints after
quantizing. In this ‘constraint quantization’ approach, due to Dirac, one treats the
constraints themselves as operators A, and demands that “physical” states ψ be those
which are solutions to the resulting equations A ψ = 0. The problem of time is associated
with the super-Hamiltonian constraint. The super-Hamiltonian H is responsible for
describing time-evolution in the classical theory, yet its counterpart in the constraint-
quantized theory, H ψ = 0, would prima facie seem to indicate that the true physical
states of the system do not evolve at all. Trying to understand how, and in what sense, the
quantum theory describes the time-evolution of something, be it states or observables, is
the essence of the problem of time.

4.2.3 Ashtekar, loop, and other variables

In geometrodynamics, all of the constraint equations are difficult to solve (though the
super-Hamiltonian constraint, known as the Wheeler-DeWitt equation, is especially
difficult), even in the absence of particular boundary conditions. Lacking solutions, one
does not have a grip on what the true, physical states of the theory are, and one cannot
hope to make much in the way of predictions. The difficulties associated with geometric
variables are addressed by the program initiated by Ashtekar and developed by his
collaborators (for a review and further references see Rovelli 2001b (Other Internet
Resources), 2001a, 2004). Ashtekar used a different set of variables, a complexified
‘connection’ (rather than a three-metric) and its canonical conjugate, which made it
simpler to solve the constraints. The program underwent further refinements with the
introduction of the loop transform, and further refinements still when it was understood
that equivalence classes of loops could be identified with spin networks. (See Smolin
(2001, 2004) for a popular introduction.)

4.3 Other Approaches

There are many other approaches to quantum gravity as well. Some (e.g., Huggett 2001,
Wüthrich 2004 (Other Internet Resources section); J. Mattingly 2005) have argued that
semiclassical gravity, a theory in which matter is quantized but spacetime is classical, is a
viable alternative. Other approaches include twistor theory (currently enjoying a revival
in conjunction with string theory), Bohmian approaches (Goldstein & Teufel, 2001),
causal sets (see Sorkin 2003, in the Other Internet Resources section) in which the
universe is described as a set of discrete events along with a stipulation of their causal
relations, and other discrete approaches (see Loll, 1998). Also of interest are arguments to
the effect that gravity itself may play a role in quantum state reduction (Christian, 2001;
Penrose, 2001).
5. Philosophical Issues
Quantum gravity raises a number of difficult philosophical questions. To date, it is the
ontological aspects of quantum gravity that have attracted the most interest from
philosophers, and it is these we will discuss in the first three sections below. In the final
section, though, we will briefly discuss the methodological and epistemological issues
which arise.

First, however, let us discuss the extent to which ontological questions are tied to a
particular theoretical framework. In its current stage of development, string theory
unfortunately provides little indication of the more fundamental nature of space, time,
and matter. Despite the consideration of ever more exotic objects — strings, p-branes, D-
branes, etc. — these objects are still understood as propagating in a background
spacetime. Since string theory is supposed to describe the emergence of classical
spacetime from some underlying quantum structure, these objects are not to be regarded
as truly fundamental. Rather, their status in string theory is analogous to the status of
particles in quantum field theory (Witten, 2001), which is to say that they are relevant
descriptions of the fundamental physics only in situations in which there is a background
spacetime with appropriate symmetries.

The duality relations between the various string theories suggest that they are all
perturbative expansions of some more fundamental, non-perturbative theory known as
‘M-theory’ (Polchinski, 2002, see the Other Internet Resources section below). This,
presumably, is the most fundamental level, and understanding the theoretical framework
at that level is central to understanding the underlying ontology of the theory. ‘Matrix
theory’ is an attempt to do just this, to provide a mathematical formulation of M-theory,
but it remains highly speculative. Thus although string theory purports to be a
fundamental theory, the ontological implications of the theory are still obscure.

Canonical quantum gravity, in its loop formulation or otherwise, has to date been of
greater interest to philosophers because it appears to confront fundamental questions in a
way that string theory, at least in its perturbative guise, does not. Whereas perturbative
string theory treats spacetime in an essentially classical way, canonical quantum gravity
treats it as quantum-mechanical, at least to the extent of treating the geometric structure
(as opposed to, say, the topological or differential structure) as quantum-mechanical.

5.1 Time

As noted in Section 4.2.2 above, the treatment of time presents special difficulties in
canonical quantum gravity. These difficulties are connected with the special role time
plays in physics, and in quantum theory in particular. Physical laws are, in general, laws
of motion, of change from one time to another. They represent change in the form of
differential equations for the evolution of, as the case may be, classical or quantum states;
the state represents the way the system is at some time, and the laws allow one to predict
how it will be in the future (or retrodict how it was in the past).
It is not surprising, then, that a theory of quantum spacetime would have a problem of
time, because there is no classical time against which to evolve the “state”. The problem
is not so much that the spacetime is dynamical; there is no problem of time in classical
general relativity. Rather, the problem is roughly that in quantizing the structure of
spacetime itself, the notion of a quantum state, representing the structure of spacetime at
some instant, and the notion of the evolution of the state, do not get any traction, since
there are no real “instants”. (In some approaches to canonical gravity, one fixes a time
before quantizing, and quantizes the spatial portions of the metric only. This approach is
not without its problems, however; see Isham (1993) for discussion and further
references.)

One can ask whether the problem of time arising from the canonical program tells us
something deep and important about the nature of time. Julian Barbour (2001a,b), for
one, thinks that it tells us that time is illusory (see also Earman (2002) in this connection).
It is argued that the fact that quantum states do not evolve under the super-Hamiltonian
means that there is no change. However, it can also be argued (Weinstein, 1999a,b) that
the super-Hamiltonian itself should not be expected to generate time-evolution; rather,
one or more “true” Hamiltonians should play this role. (See Butterfield & Isham (1999)
and Rovelli (2006) for further discussion.)

5.2 Ontology

The problem of time is closely connected with a general puzzle about the ontology
associated with “quantum spacetime”. Quantum theory in general resists any
straightforward ontological reading, and this goes double for quantum gravity. In
quantum mechanics, one has particles, albeit with indefinite properties. In quantum field
theory, one again has particles (at least in suitably symmetric spacetimes), but these are
secondary to the fields, which again are things, albeit with indefinite properties. On the
face of it, the only difference in quantum gravity is that spacetime itself becomes a kind
of quantum field, and one would perhaps be inclined to say that the properties of
spacetime become indefinite. But space and time traditionally play important roles in
individuating objects and their properties—in fact a field is in some sense a set of
properties of spacetime points — and so the quantization of such raises real problems for
ontology.

In the loop quantum gravity program, the area and volume operators have discrete
spectra. Thus, like spins, they can only take certain values. This suggests (but does not
imply) that space itself has a discrete nature, and perhaps time as well (depending on how
one resolves the problem of time). This in turn suggests that space does not have the
structure of a differential manifold, but rather that it only approximates such a manifold
on large scales, or at low energies.

5.3 Status of quantum theory

Whether or not spacetime is discrete, the quantization of spacetime entails that our
ordinary notion of the physical world, that of matter distributed in space and time, is at
best an approximation. This in turn implies that ordinary quantum theory, in which one
calculates probabilities for events to occur in a given world, is inadequate as a
fundamental theory. As suggested in the Introduction, this may present us with a vicious
circle. At the very least, one must almost certainly generalize the framework of quantum
theory. This is an important driving force behind much of the effort in quantum
cosmology to provide a well-defined version of the many-worlds or relative-state
interpretations. Much work in this area has adopted the so-called ‘decoherent histories’ or
‘consistent histories’ formalism, whereby quantum theories are understood to make
probabilistic predictions about entire (coarse-grained) ‘histories’. Almost all of this work
to date construes histories to be histories of spatiotemporal events, and thus presupposes a
background spacetime; however, the incorporation of a dynamical, quantized spacetime
clearly drives much of the cosmology-inspired work in this area.

More generally, one might step outside the framework of canonical, loop quantum
gravity, and ask why one should only quantize the metric. As pointed out by Isham
(1994, 2002), it may well be that the extension of quantum theory to general relativity
requires one to quantize, in some sense, not only the metric but also the underlying
differential structure and topology. This is somewhat unnatural from the standpoint where
one begins with classical, canonical general relativity and proceeds to “quantize” (since
the topological structure, unlike the metric structure, is not represented by a classical
variable). But one might well think that one should start with the more fundamental,
quantum theory, and then investigate under which circumstances one gets something that
looks like a classical spacetime.

5.4 Methodology

The nature of the enterprise, in particular its seeming remoteness from experiment, gives
rise to significant methodological and epistemological questions as well, focusing on the
problem of how to construct or discover a scientific theory for phenomena which are so
remote from observation. Are beauty and consistency either necessary or sufficient? The
pronounced split between the string theory community and the loop quantum gravity
community has nothing to do with empirical success or lack thereof, but much to do with
factors which might normally play a role only on the periphery of the scientific
enterprise. The history of the enterprise of quantum gravity might well be worth
historico-philosophical scrutiny, much as the history of cosmology has been, cosmology
having also been rather data-starved until recently. (See Kragh (1999) for an excellent
account of the big-bang/steady-state controversy.)

6. Conclusion
In the author's opinion, it is unlikely that a final theory of quantum gravity — if indeed
there is one — will look much like any of the current candidate theories, be they string
theory, canonical gravity, or other approaches. However, the philosophical and
conceptual study of quantum gravity is useful insofar as it prompts one to consider
questions which are surely raised by this almost quixotic undertaking, questions which
will ultimately require some unknown combination of philosophical, physical, and
mathematical insight to answer.

Bibliography
• Baez, J., 2001, “Higher-dimensional algebra and Planck scale physics,” in
Callender & Huggett, 177-195, [Preprint available online].
• Barbour, J., 2001a, “On general covariance and best matching,” in Callender &
Huggett, 199-212.
• -----, 2001b, The End of Time: The Next Revolution in Physics, Oxford University
Press, Oxford.
• Bekenstein, J., 1973, “Black holes and entropy”, Physical Review D 7, 2333-
2346.
• Belot, G. and Earman, J., 2001, “Pre-Socratic quantum gravity,” in Callender &
Huggett, 213-255.
• Brown, H. and Pooley, O., 2001, “The origin of the spacetime metric: Bell's
‘Lorentzian pedagogy’ and its significance in general relativity,” in Callender &
Huggett, 256-272, [Preprint available online].
• Butterfield, J. and Isham, C., 1999, “On the emergence of time in quantum
gravity” in The Arguments of Time, ed. by J. Butterfield, British Academy and
Oxford Univ. Press, 111-168, [Preprint available online].
• -----, 2001, “Spacetime and the philosophical challenge of quantum gravity” in
Callender & Huggett, 33-89, [Preprint available online].
• Callender, C. and Huggett, N., 2001, Physics Meets Philosophy at the Planck
Scale, Cambridge University Press, Cambridge.
• Carlip, S., 2001,“Quantum gravity: a progress report”, Reports on Progress in
Physics 64: 885-942, [Preprint available online].
• Cao, T.Y., 2001, “Prerequisites for a consistent framework of quantum gravity,”
Studies in the History and Philosophy of Modern Physics 32B, 181-204.
• Christian, J., 2001, “Why the quantum must yield to gravity,” in Callender &
Huggett, 305-338, [Preprint available online].
• Earman, J., 2002, “A thoroughly modern McTaggart. Or what McTaggart would
have said if he had learned general relativity theory,” Philosophers' Imprint 2/3
[Available online in PDF from the publisher
• Goldstein, S. and Teufel, S., 2001, “Quantum spacetime without observers:
Ontological clarity and the conceptual foundations of quantum gravity,” in
Callender & Huggett, 275-289, [Preprint available online].
• Greene, B., 2000, The Elegant Universe, Vintage.
• Hawking, S., 1974, “Black hole explosions,” Nature 248, 30-31.
• Huggett, N., 2001, “Why quantize gravity (or any other field for that matter),”
Philosophy of Science, 68: (3), S382-S394.
• Isham, C.J., 1993, “Canonical quantum gravity and the problem of time,” in
Integrable Systems, Quantum Groups, and Quantum Field Theories, L.A. Ibort
and M.A. Rodriguez (eds.), Kluwer, Dordrecht, 157-288, [Preprint available
online].
• -----, 1994, “Prima facie questions in quantum gravity,” in Canonical Gravity:
From Classical to Quantum (Lecture Notes in Physics 434), J. Ehlers and H.
Friedrich (eds.), Springer-Verlag, Berlin, 1-21, [Preprint available online].
• -----, 2002, “Some reflections on the status of conventional quantum theory when
applied to quantum gravity,” in The Future of Theoretical Physics and
Cosmology, G. Gibbons, E. Shellard, and S. Rankin (eds.), Cambridge University
Press, Cambridge, 384-408, [Preprint available online].
• Kragh, H., 1999, Cosmology and Controversy, Princeton Univ. Press, Princeton,
1999.
• Kuchař, K., 1992, “Time and interpretations of quantum gravity,” in Proceedings
of the 4th Canadian Conference on General Relativity and Astrophysics, edited by
G. Kunstatter, D. Vincent, and J. Williams, Singapore: World Scientific, Preprint
available online].
• -----, 1993, “Canonical quantum gravity,” in General Relativity and Gravitation
1992: Proceedings of the Thirteenth International Conference on General
Relativity and Gravitation, edited by R. Gleiser, C. Kozameh and O. Moreschi,
IOP Publishing, Bristol, [Preprint available online].
• Loll, R., 1998, “Discrete approaches to quantum gravity in four dimensions,”
Living Reviews in Relativity 1/13 (version cited = lrr-1998-13) [Available online
from the publisher].
• Mattingly, D., 2005, “Modern tests of Lorentz invariance,” Living Reviews in
Relativity 8, No. 5 (cited on October 15, 2005) (version cited = lrr-2005-5)
[Available online from the publisher].
• Mattingly, J., 2005, “Is quantum gravity necessary?” in The Universe of General
Relativity: Einstein Studies, volume 11, Jean Eisenstaedt, Anne Kox (eds.),
Birkhäuser, Boston.
• Penrose, R., 2001, “On gravity's role in quantum state reduction,” in Callender &
Huggett, 290-304.
• Rickles, D., 2005, “A new spin on the hole argument,”, Studies in the History and
Philosophy of Modern Physics 36, 415-434, [Preprint available online].
• -----, 2006, “Time and structure in canonical gravity,” in The Structural
Foundations of Quantum Gravity, ed. by D. Rickles, S. French, and J. Saatsi,
Clarendon Press, Oxford, [Preprint available online].
• Rovelli, C., 2001a, “Quantum spacetime: What do we know?,” in Callender &
Huggett, 101-122.
• -----, 2002, “Partial observables,” Physical Review D 65, 124013, [Preprint
available online].
• -----, 2004, Quantum Gravity, Cambridge University Press, Cambridge.
• -----, 2006 (forthcoming), “Quantum gravity,” in The Handbook of Philosophy of
Physics, ed. by J. Butterfield and J. Earman, North Holland.
• Smolin, L. 2001, Three Roads to Quantum Gravity, Basic Books.
• -----, 2004, “Atoms of space and time,” Scientific American, January, 66-75.
• Sorkin, R., 1997, “Forks in the road, on the way to quantum gravity,”
International Journal of Theoretical Physics, 36, 2759-2781, [Preprint available
online].
• Unruh, W., 2001, “Black holes, dumb holes, and entropy,” in Callender &
Huggett, 152-173.
• Weingard, R., 2001, “A philosopher looks at string theory,” in Callender &
Huggett, 138-151.
• Weinstein, S., 1999a, “Gravity and gauge theory,” Philosophy of Science 66,
S146-S155, [Preprint available online].
• -----, 1999b, “Time, gauge, and the superposition principle in quantum gravity,”
in The Eighth Marcel Grossmann Meeting on General Relativity, Tsvi Piran (ed.),
World Scientific, Singapore, [Preprint available online].
• -----, 2001a, “Absolute quantum mechanics” British Journal for the Philosophy of
Science 52, 67-73, [Preprint available online].
• -----, 2001b, “Naive quantum gravity,” in Callender & Huggett, 90-100, [Preprint
available online].
• Witten, E., 2001, “Reflections on the fate of spacetime,” in Callender & Huggett,
125-137.

Other Internet Resources


Unpublished, Online Manuscripts

• Amelino-Camelia, G., 2003, “Quantum gravity phenomenology”.


• Horowitz, G., 2000, “Quantum gravity at the turn of the millenium”.
• Polchinski, J., 2002, “M-theory: uncertainty and unification”.
• Rovelli, C., 2001b, “Notes for a brief history of quantum gravity”, (Centre De
Physique Théorique, University of Marseilles).
• Smolin, L., 2003, “How far are we from the quantum theory of gravity?”.
• Susskind, L., 2003, “The anthropic landscape of string theory”.
• Sorkin, R., 2003, “Causal sets: discrete gravity”, (Notes for the Valdivia Summer
School).
• Torre, C., 2005, “Observables for the polarized Gowdy model”.
• Wüthrich, C., 2004, “To quantize or not to quantize: fact and folklore in quantum
gravity”.

You might also like