quantum gravitational optics

20

Click here to load reader

Upload: graham-m

Post on 08-Dec-2016

220 views

Category:

Documents


4 download

TRANSCRIPT

Page 1: Quantum gravitational optics

This article was downloaded by: [University of Western Ontario]On: 12 May 2013, At: 07:25Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954 Registered office: MortimerHouse, 37-41 Mortimer Street, London W1T 3JH, UK

Contemporary PhysicsPublication details, including instructions for authors and subscription information:http://www.tandfonline.com/loi/tcph20

Quantum gravitational opticsGraham M Shorea Department of Physics, University of Wales, Swansea, UKPublished online: 25 May 2010.

To cite this article: Graham M Shore (2003): Quantum gravitational optics , Contemporary Physics, 44:6, 503-521

To link to this article: http://dx.doi.org/10.1080/00107510310001617106

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

This article may be used for research, teaching, and private study purposes. Any substantial or systematicreproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form toanyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contentswill be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug dosesshould be independently verified with primary sources. The publisher shall not be liable for any loss, actions,claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly orindirectly in connection with or arising out of the use of this material.

Page 2: Quantum gravitational optics

Quantum gravitational optics

GRAHAM M. SHORE

In quantum theory, the curved spacetime of Einstein’s general theory of relativity acts as a

dispersive optical medium for the propagation of light. Gravitational rainbows and

birefringence replace the classical picture of light rays mapping out the null geodesics of

curved spacetime. Even more remarkably, superluminal propagation becomes a real

possibility, raising the question of whether it is possible to send signals into the past. In this

article, we review recent developments in the quantum theory of light propagation in general

relativity and discuss whether superluminal light is compatible with causality.

1. Time and light

Einstein’s discovery of relativity nearly a century ago

highlighted the fundamental nature of the speed of light

and revolutionised our concept of time. In particular,

relativity predicts that a signal moving faster than light

would, from the viewpoint of some inertial observers, be

moving backwards in time. The ideas of faster than light

motion and time travel became inextricably linked.

Recently, a new perspective on these issues has arisen from

the study of quantum effects such as vacuum polarisation

on the nature of light propagation in general relativity. The

classical picture of light rays mapping out the geometry of

curved spacetime gives way in quantum theory to a range

of optical phenomena such as birefringence, lensing and

dispersion. Gravitational rainbows appear. Most remark-

ably, even superluminal propagation appears to be

possible.

In this paper, we review some of these recent develop-

ments, focusing on the issue of superluminal light and its

implications for causality. From the outset, we have to

clarify what we mean by ‘superluminal’. Special (general)

relativity is characterised by global (local) Lorentz invar-

iance, that is the invariance of the laws of physics in

different inertial frames of reference. The Lorentz trans-

formations introduce a new fundamental constant of

nature, c, which in classical theory may be identified with

the unique and unambiguous speed of light. In what

follows, however, we must distinguish clearly between this

fundamental constant and the actual propagation speed of

light; when we use the word ‘superluminal’, we simply mean

‘greater than c’. The key point is that it is the null cones

generated by the speed c which divide spacetime into

‘timelike’ and ‘spacelike’ regions; a superluminal signal

therefore travels into the spacelike region and, for some

observers, back in time. As we shall see, light may do

precisely this.

The propagation of light is already a rich field in classical

general relativity. In the presence of a gravitational field,

spacetime becomes curved. Light rays follow null geodesics,

geodesics being the analogues of shortest distant paths in

curved spacetime, and null denoting that the frequency and

wave vector satisfy the simple relation k2 ¼ o2 � jkj2 ¼ 0

(where from now on, we use conventional units where

c=1) which is equivalent to the phase velocity

uph ¼ o=jkj ¼ 1. Light rays therefore act as tracers,

mapping out the curvature of spacetime. This immediately

predicts the bending of light around a massive object, as

observed by Eddington during the famous 1919 expedition

to observe a solar eclipse in Brazil. Nowadays, this effect is

the basis for the burgeoning science of gravitational

lensing, which has become a crucial tool in the search for

dark matter in the universe.

In quantum field theory, however, a whole range of new

phenomena emerge. Curved spacetime acts as an optical

medium with its own refractive index. Dispersive effects

(rainbows) and polarisation-dependent propagation (bire-

fringence) arise. The simple identification of the geometric

null cones of curved spacetime with the physical light cones

corresponding to the actual propagation of light pulses is

lost. The field of quantum gravitational optics takes on a

new richness. The big surprise, however, is the apparentAuthor’s address: Department of Physics, University of Wales, Swansea,UK. E-mail: [email protected]

Contemporary Physics, November –December 2003, volume 44, number 6, pages 503 – 521

Contemporary Physics ISSN 0010-7514 print/ISSN 1366-5812 online # 2003 Taylor & Francis Ltdhttp://www.tandf.co.uk/journals

DOI: 10.1080/00107510310001617106

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 3: Quantum gravitational optics

prediction in some circumstances of a superluminal speed

of light.

These new developments were initiated by a pioneering

paper in 1980 by Drummond and Hathrell [1], who studied

the effects on light propagation of vacuum polarisation (see

figure 1) in quantum electrodynamics. This is a uniquely

quantum field theoretic process, in which a photon

metamorphoses into a virtual electron-positron pair. This

gives the photon an effective size characterised by the

Compton wavelength lc of the electron. It follows that its

propagation will be affected by gravity if the scale of the

spacetime curvature is comparable to lc. Of course, such

effects will be tiny for the curvatures associated with

normal astrophysical objects. However, this is character-

istic of the physics of quantum fields in curved spacetime,

the most important example of which is the famous

Hawking [2] radiation from black holes. Like Hawking

radiation, the main importance of the effects described here

is likely to be conceptual, especially on their implications

for our understanding of the realisation of causality in

theories incorporating quantum theory and gravity.

As Drummond and Hathrell showed, the effect of

vacuum polarisation is to induce direct interactions of

O(a) between the electromagnetic field and the spacetime

curvature. In other words, the classical Maxwell’s equa-

tions are modified by terms explicitly involving the

curvature. This breaks the strong equivalence principle

(SEP). As we explain carefully later, it is this effective

violation of the SEP which opens up the possibility that

superluminal propagation may exist without, as would be

the case in special relativity, necessarily implying a break-

down of causality. This is the key conceptual issue raised by

the introduction of quantum field theory into gravitational

optics.

The paper is arranged as follows. In section 2, we

present an introduction to the mathematics of special and

general relativity for readers who are not entirely familiar

with the concepts and notation used in the paper. Special

emphasis is placed on the generalisation of Maxwell’s

equations of electrodynamics to a general relativistic

setting. In section 3, we review briefly the key points of

geometric optics, which allow us to pass from Maxwell’s

equations, or their quantum modifications, to the char-

acteristics of light propagation. Vacuum polarisation in

curved spacetime and the construction of the Drummond-

Hathrell effective action are reviewed in section 4, and

their implications for photon propagation in a variety of

cosmological and black hole spacetimes are explored in

section 5. Section 6 is devoted to the conceptual issues

surrounding the compatibility of superluminal light with

causality. Section 7 focuses on the crucial question of

dispersion, explaining why the high-frequency limit is

essential for causality and describing recent developments

[3, 4, 5] in quantum theory, including a generalised

effective action aimed at clarifying the nature of high-

frequency propagation. Finally, in section 8, we comment

briefly on a number of speculative ideas in the current

literature which reflect in some way on the physics

reviewed here.

2. A primer on electrodynamics in curved spacetime

In this section, we introduce some elements of the

geometrical description of electrodynamics in special and

general relativity that we will use throughout the paper.

Special relativity implies that events can be considered as

taking place in the arena of a 4-dimensional spacetime,

known as Minkowski spacetime. This differs from a

Euclidean space in that the metric, i.e. the distance between

two neighbouring points (t, x, y, z) and (t+ dt, x+ dx,

y+ dy, z+ dz), is given by

ds2 ¼ c2dt2 � dx2 � dy2 � dz2: ð1Þ

The crucial sign differences distinguish ‘time’ from

‘space’ coordinates. The metric therefore divides space-

time intervals into 3 types: ‘timelike’, for which ds24 0;

‘spacelike’, for which ds25 0; and ‘lightlike’ or ‘null’,

for which ds2=0. If we consider a particle moving

along the x-axis with ds2=0, then clearly its velocity is

u= dx/dt= c, so that the fundamental parameter enter-

ing the Minkowski metric is identified as the speed of

light. Light (photons) therefore travel along null paths.

Massive particles on the other hand travel with speeds

less than c, i.e. along timelike paths. Normally, it is

assumed that nothing can travel along spacelike paths

since this would require a superluminal velocity u4 c.

This possibility, however, is one of the main topics of

this review!

In this Minkowski spacetime geometry, the objects which

take over the role of the familiar vectors of 3-dim Euclidean

space are 4-vectors. The simplest example is the spacetime

position 4-vector xm=(t, x), which has time and space

components distinguished by the index m=0 (time) or

m=1, 2, 3 (space). In this notation, the metric (1) is written

as

Figure 1. The O(a) vacuum polarisation Feynman diagram

contributing to the photon propagator (wavy line). The solid lines

represent the electron (positron) propagator in curved spacetime.

504 G. M. Shore

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 4: Quantum gravitational optics

ds2 ¼ gmndxmdxn; ð2Þ

where

gmn ¼c2 0 0 00 �1 0 00 0 �1 00 0 0 �1

0BB@

1CCA

is known as the metric tensor. A summation convention

over repeated indices is assumed.1

Under a change of frame of reference, or in other words

viewed from the perspective of a relatively moving

observer, the spacetime coordinates of a given event

change. This change is known as a Lorentz transformation.

Viewed in a frame of reference moving along the x-axis

with velocity u, the spacetime point xm=(t, x, y, z) is

described as x’m=(t’, x’, y’, z’), where

t0 ¼ gðuÞ�t � u

c2x�;

x0 ¼ gðuÞ�x � ut�; ð3Þ

with gðuÞ ¼ ð1� u2

c2Þ�1

2, y’= y and z’= z. A crucial point is

that the metric ds2 is invariant under Lorentz transforma-

tions. This means that the characterisation of a given path

as timelike, spacelike, or null is independent of the frame of

reference. In particular, all observers agree whether a path

is null, i.e. they all observe the same speed of light.

In order to formulate electrodynamics in Minkowski

spacetime, we need to find an appropriate geometrical

description of the electric and magnetic fields E and B.

Since these involve 6 component fields, clearly they cannot

fit into a 4-vector. In fact, the resolution is to write them as

the independent components of an antisymmetric 46 4

matrix or tensor F mn, viz:

F mn ¼0 1

c Ex1c Ey

1c Ez

� 1c Ex 0 Bz �By

� 1c Ey

�Bz 0 Bx

� 1c Ez

By �Bx 0

0BBB@

1CCCA: ð4Þ

The Lorentz transformation properties of F mn follow

straightforwardly as a generalisation of those for the

spacetime vector. That is, if we represent (3) by

x0m ¼ Lmnx

n; ð5Þ

then

F 0mn ¼ LmlL

nrF

lr: ð6Þ

It is straightforward to check that this geometric form of

the Lorentz transformation reproduces the standard result

for the transformation of the electric and magnetic fields

observed in different frames of reference:

E 0x ¼ Ex;E

0y ¼ gðuÞ�Ey � uBz

�;E 0

z ¼ gðuÞ�Ez þ uBy

�;

B 0x ¼ Bx;B

0y ¼ gðuÞ�By þ u

c2Ez

�;B 0

z ¼ gðuÞ�Bz � u

c2Ey

�: ð7Þ

The next step is to transcribe Maxwell’s equations into

geometric form. (We shall do this here just for the source-

free equations. The generalisation may be found in any

good relativity or electromagnetism textbook). Recall these

fall into two pairs:

r:B ¼ 0;

r�Eþ @B

@t¼ 0; ð8Þ

and

r:E ¼ 0;

r�B� 1

c2@E

@t¼ 0; ð9Þ

where c ¼ 1=ffiffiffiffiffiffiffiffiffim0E0

p. Only the second pair is modified by the

inclusion of sources. The first pair constrains the form of

the fields and implies the existence of electric and magnetic

potentials f and A such that

E ¼ �rf� @A

@t;

B ¼ r�A: ð10Þ

Now, in the geometric formalism the potentials can be

combined into an electromagnetic potential 4-vector

Am=(f, A) with the appropriate Lorentz transformation

A’m=LmnA

n. Equations (8) become

@lF mn þ @mF nl þ @nF lm ¼ 0; ð11Þ

which imply

F mn ¼ @mAn � @nAm; ð12Þ

where we have used the simplified notation @m for @/@xm. It

is then easy to check that the remaining Maxwell’s

equations (9) can be written geometrically as the single

equation

@mFmn ¼ 0: ð13Þ

Evidently, this geometrical expression of the Maxwell’s

equations in terms of spacetime 4-vectors and tensors is very

elegant. It is much more than that, however. It demon-

strates that Maxwell’s equations themselves are already

fully compatible with special relativity, or in technical

terms, they are Lorentz invariant. Unlike Newtonian

dynamics, Maxwell’s electromagnetism does not need any

modification to satisfy the principles of special relativity.

1In Minkowski spacetime, it is also necessary to distinguish betweencontravariant 4-vectors such as xm and the equivalent covariant 4-vectors xmdefined by contraction with the metric tensor, i.e. xm=gmnx

n. Raising indicesis performed by contracting with gmn, the inverse of the metric tensor gmn.

505Quantum gravitational optics

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 5: Quantum gravitational optics

This simple picture of electrodynamics in special

relativity has to be modified in the presence of gravity,

i.e. in general relativity. As is well-known, the presence of a

gravitational field is reflected in a change in the nature of

spacetime, instead of being simply ‘flat’, Minkowskian

spacetime is curved. Roughly speaking, particles travel

along shortest-distance paths through spacetime, known as

geodesics [see (25)], and the gravitational force is mani-

fested by a change in the separation of two particles as they

each follow a different geodesic path.

The key to general relativity is the weak equivalence

principle, which asserts that the force of gravity disappears

in a freely-falling frame of reference. Mathematically, this

states that at each point in spacetime, there is a local frame

of reference in which spacetime appears Minkowskian (a

so-called Lorentz frame), i.e. the curved spacetime must be

locally flat. This turns out to imply that the metric must

take the form

ds2 ¼ gmndxmdxn; ð14Þ

as before, but where now gmn is no longer the simple

Minkowski metric tensor but has a more interesting

functional formwhichencodes thecurvatureofthespacetime.

Several examples of metrics for the spacetimes describing

black holes, gravitational waves, cosmology, etc. will appear

throughout this review. Curved spacetimes with this special

property are known as Riemannian. General relativity is

therefore the theory of Riemannian curved spacetimes.

An extension of the weak equivalence principle, and one

we will be forced to question later, states that the laws of

physics in curved spacetime are such that they reduce to the

usual special relativistic form in the local Lorentz frame of

reference at each spacetime point. This rule (known as the

strong equivalence principle) specifies how, for example,

electrodynamics is to be translated from Minkowski to

Riemannian spacetime.

Essentially, all that is required is to replace the ordinary

derivatives @m appearing in any equation by so-called

covariant derivatives Dm, which are constructed to compen-

sate for the spacetime curvature. For example, the

definition of the field strength tensor Fmn can be rewritten as

F mn ¼ DmAn �DnAm; ð15Þ

where the covariant derivative Dm acting on the 4-vector An

is defined by

DmAn ¼ @mA

n þ �nmlA

l: ð16Þ

The Christoffel symbols Gnmlare constructed from the metric

tensor:

�nml ¼

1

2gnr�@mgrl þ @lgmr � @rgml

�: ð17Þ

Similarly, Maxwell’s equation (13) generalises simply to

DmFmn ¼ 0: ð18Þ

This concludes our brief introduction to the geometric

description of electrodynamics in special and general

relativity. In the rest of the paper, we go on to use this

formalism to study the propagation of light and in

particular, how the inclusion of quantum effects changes

this simple picture in remarkable ways, notably in raising

the possibility that light may travel faster than the

fundamental ‘speed of light’ parameter c.

3. Geometric optics

As we have seen, the propagation of light in general

relativity is governed classically by Maxwell’s equation for

the electromagnetic field,

Dm F mn ¼ 0; ð19Þ

where Fmn=DmAn –DnAm. The simplest way to deduce

the properties of light cones and light rays from this

equation is to use geometric optics. (For a more detailed

explanation, see for example reference [6].) This starts

from the ansatz

ðAm þ iEBm þ . . .Þexp�i#E

�; ð20Þ

in which the electromagnetic field is written as a slowly-

varying amplitude and a rapidly-varying phase. The

parameter E is introduced purely as a book-keeping device

to keep track of the relative order of magnitude of terms,

and the Maxwell’s equations are solved order-by-order in E.The idea here is that we consider waves where the

wavelength is much smaller than the scale on which either

the amplitude or the background spacetime curvature is

varying. This means that it is sensible to think of the wave

locally as being just like a standard, undistorted plane

wave.

The wave vector is identified as the gradient of the

phase, km= @m#. To see why, remember that in the usual

Minkowski spacetime, a plane wave propagating in the z-

direction would be represented by an amplitude times the

phase factor exp[i(ot – kz)]. Differentiating the phase

w.r.t. xm as above just gives the 4-dim wave vector

km=(o, 0, 0, k). The corresponding phase velocity of the

wave is uph=o/k. We also write Am=Aam, where A

represents the amplitude itself while am specifies the

polarisation. Since the physical polarisations are trans-

verse to the direction of motion, am satisfies kmam=0. We

also normalise it so that amam=–1 (since it is a spacelike

4-vector).

Solving Maxwell’s equation, we find at O(1/(E),

506 G. M. Shore

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 6: Quantum gravitational optics

k2 ¼ 0; ð21Þ

while at O(1),

kmDman ¼ 0; ð22Þ

and

kmDmðlnAÞ ¼ � 1

2Dmk

m: ð23Þ

Equation (21) shows immediately that km is a null vector.

From its definition as @m# and the symmetry of Christoffel

symbols, it follows immediately that Dmkn=Dnkm and so

kmDmkn ¼ kmDnkm ¼ 1

2Dnk2 ¼ 0: ð24Þ

Light rays, or equivalently photon trajectories, are the

curves whose tangent is km, i.e. the curves xm(s) where dxm/

ds= km. These curves therefore satisfy

0 ¼ kmDmkn ¼ d 2xn

ds2þ Gn

mldxm

ds

dxl

ds: ð25Þ

This is the geodesic equation. We conclude that for the

usual Maxwell theory in general relativity, light follows

null geodesics. Equations (24) and (22) show that both

the wave vector and the polarisation are parallel

transported along these null geodesic rays, while equa-

tion(23) shows how the amplitude changes as the beam

of rays focuses or diverges.

This picture allows us to identify the physical light cones,

specified by the relation k2=0, with the geometric null

cones of the background spacetime. Moreover, light

propagation is non-dispersive, i.e. light of all frequencies

(or equivalently, photons of all momenta), travel at the

same unique speed c, the fundamental constant determining

the local Lorentz invariance of curved spacetime. Of

course, this is why the distinction between ‘timelike’ and

‘spacelike’ directions and more generally the analysis of the

causal properties of curved spacetime is traditionally

expressed in the language of light signals. However, as we

shall now see, this whole picture has to be radically revised

when we consider the quantum nature of light and its

interaction with gravity.

4. Vacuum polarisation

The quantum theory of photons interacting with electrons

is quantum electrodynamics, so in this section we take a

first look at the effects of QED interactions on photon

propagation in curved spacetime, i.e. in a classical back-

ground gravitational field.

The simplest Feynman diagram contributing to photon

propagation is the so-called ‘vacuum polarisation’ process

shown in figure 1. In picturesque terms, this can be thought

of as the photon splitting into a virtual electron-positron

pair which subsequently recombine. The effect is, in a

certain sense, to give the photon an effective size

characterised by the Compton wavelength lc=1/m of the

electron (where m is its mass). It is therefore possible to

think of the photon propagating as if it were a quantum

cloud of size O(lc) rather than a point particle. It is then

plausible that if the photon passes through an anisotropic

curved spacetime whose typical curvature scale L is

comparable to lc, then its motion would be affected,

possibly in a polarisation-dependent way.

A rigorous analysis in QED confirms this picture. The

light-cone is indeed altered and the photon acquires a

polarisation-dependent correction to its velocity of

Oðal2c=L2Þ. What is less predictable is that in some cases

these quantum corrections apparently produce superlum-

inal velocities!

In quantum field theory, the effect of all the quantum

processes such as vacuum polarisation can be neatly

summarised by writing a new effective action, which

modifies the usual Maxwell action

GM ¼ � 1

4

Zdx

ffiffiffig

pFmnF

mn: ð26Þ

The calculation required to translate Feynman diagrams

such as figure 1 into an effective action involves advanced

techniques in QED and will not be presented here.

However, once we have the new effective action, we can

derive equations of motion in the usual way. These

equations will then be the modifications of Maxwell’s

equations which encode the effect of vacuum polarisation

on the propagation of light in curved spacetime.

The simplest approximation to the effective action was

first calculated by Drummond and Hathrell in their

pioneering paper on superluminal light [1]. This is:

GDH ¼ GMþ 1

m2

Zdx

ffiffiffiffiffiffiffi�gp �

aRFmnFmn þ bRmnF

mlF nlþ

cRmnlrFmnF lr þ dDmF

mlDnFnl

�; ð27Þ

where a, b, c, d are perturbative coefficients of O(a), viz.

a ¼ � 1

144

ap; b ¼ 13

360

ap; c ¼ � 1

360

ap; d ¼ � 1

30

ap: ð30Þ

To understand this, notice first that all the new terms

have coefficients proportional to the fine structure constant

a (which is approximately 1/137). This shows that they have

come from so-called ‘one-loop’ Feynman diagrams of the

type shown in figure 1. More complicated diagrams with

more loops give terms of higher order in a but these are of

course much smaller. Next, they depend on the spacetime

507Quantum gravitational optics

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 7: Quantum gravitational optics

curvature, represented by one of the curvature tensors R,

Rmn or Rmnlr (the Ricci scalar, Ricci tensor and Riemann

tensor respectively). As we emphasise in section 6, this is the

radically new effect of vacuum polarisation, since interac-

tions like this would be forbidden by the strong equivalence

principle and therefore not normally included in a classical

theory of electrodynamics in curved spacetime. What we

find here is that they inevitably appear in the quantum

theory.

In fact, even at O(a), a complete evaluation of the

effective action is impossible with current techniques. In

writing the Drummond-Hathrell action, we were therefore

forced to make two crucial approximations. The first is a

weak-field approximation for gravity. This means we keep

only terms in the final effective action of first order in the

curvature tensors R, Rmn and Rmnlr. Applying simple

dimensional analysis, we see that these terms have

dimensions (in �h= c=1 units) of (mass)2 and so must be

balanced by a compensating 1/m2 parameter. The only such

mass scale in the vacuum polarisation Feynman diagram is

the electron mass. This implies that our results for photon

propagation are valid only to the lowest order in the

parameter ‘R’/m2*l2c/L2, where L is the typical curvature

scale. The second approximation, which we shall try to

overcome in section 7, is that we are restricted to low-

frequency photons. From the discussion of geometrical

optics above, it can be seen that this is equivalent to

neglecting terms induced in the effective action which

involve higher orders in derivatives of the fields. Fortu-

nately, even with these approximations, the resulting

effective action already encodes much of the novel physics

of photon propagation in curved spacetime.

5. Faster than light?

In this section, we show how photon propagation is

modified by these new interactions between photons and

the gravitational field. As we shall see, this brings some

surprises, notably the possibility of superluminal velocities.

5.1. Light cones and birefringence

The first step is to apply the geometric optics method

described in section 3 to the equation of motion derived

from the quantum effective action (27), viz:

DmFmn � 1

m2

2bR mlD

mF ln þ 4cR mnlrD

mF lr¼ 0: ð29Þ

Here, we have neglected derivatives of the curvature tensor,

which would be suppressed by powers of O(l/L), where l is

the photon wavelength. We have also omitted the

contribution on the r.h.s. involving the Ricci scalar, which

is proportional to 2aRDmFmn. Since DmF

mn is already O(a)using the equations of motion, this contribution only

affects the light cone condition at O(a2) and must be

dropped for self-consistency, since all the other terms on

the r.h.s. have only been evaluated up to O(a). (The purelyelectromagnetic contribution with coefficient d also does

not contribute for the same reason). Physically, this reflects

the fact that some anisotropy is required to modify the

speed of light. A constant curvature spacetime, where both

Rmn and Rmnlr can be expressed just in terms of the Ricci

scalar R and factors of the metric, is not sufficient.

Implementing the geometric optics ansatz, we quickly

find the new light cone condition:

k2 � 2b

m2Rmlk

mkl þ 8c

m2Rmnlrk

mklanar ¼ 0: ð30Þ

Since k2„ 0, this shows immediately that the speed of light

is changed, it is no longer equal to the fundamental

constant c! The gravitational field has affected the speed of

light, just like an optical medium such as a prism.

However, since equation (30) is still homogeneous and

quadratic in km, we can write it as

Gmnkmkn ¼ 0; ð31Þ

defining Gmn as the appropriate function of the curvature

and polarisation. This immediately means that, at least in

this approximation, there is no dispersion. The reason is

that if we express (31) in terms of its components in an

orthonormal frame, it typically takes the form

Ak0k0 � Bijkikj ¼ 0; ð32Þ

where A=1+O(a) and Bij= dij+O(a) are coefficients

derived from Gmn. The phase velocity uph= k0/

jkj=1+O(a) is therefore just a simple number, indepen-

dent of the frequency o:k0.2

Now notice that in the discussion of the free Maxwell’s

theory, we did not need to distinguish between the photon

momentum pm, i.e. the tangent vector to the light rays, and

the wave vector km since they were simply related by raising

the index using the spacetime metric, pm= gmnkn. In the

modified theory, however, there is an important distinction.

The wave vector, defined as the gradient of the phase, is a

covariant vector or 1-form, whereas the photon momen-

tum/tangent vector to the rays is a true contravariant

vector. The relation is non-trivial. In fact, given km, we

should define the corresponding ‘momentum’ as

2Notice also that this is quite different from the dispersion relation p2=–m2 for a ‘conventional’ tachyon. In this case (see section 6), thesuperluminal tachyon velocity is energy dependent.

508 G. M. Shore

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 8: Quantum gravitational optics

pm ¼ Gmnkn; ð33Þ

and the light rays as curves xm(s) where dxm/ds= pm. This

definition of momentum satisfies

Gmnpmpn ¼ Gmnkmkn ¼ 0; ð34Þ

where G � G�1 defines a new effective metric which

determines the light cones mapped out by the geometric

optics light rays. (Indices are always raised or lowered using

the true metric gmn.) The ray velocity uray corresponding to

the momentum pm, which is the velocity with which the

equal-phase surfaces advance, is given by (defining

components in an orthonormal frame)

uray ¼ jpjp0

¼ djxjdt

; ð35Þ

along the ray. This is in general different from the phase

velocity

uph ¼ k0

jkj ; ð36Þ

although this discrepancy only arises when there is a

difference between the direction of propagation along the

rays and the wave 3-vector. Otherwise, it follows directly

from (34) that uray and uph are identical.

This shows that, at least at this level of approximation,

photon propagation for QED in curved spacetime can be

characterised as a bimetric theory; the physical light cones

are determined by the effective metric Gmn and are distinct

from the geometric null cones which are fixed by the

spacetime metric gmn.

It is instructive to rewrite the new light cone condition in

terms of the Weyl tensor and the energy-momentum tensor,

using the Einstein field equation

Rmn � 1

2Rgmn ¼ 8pTmn: ð37Þ

Notice that here we have used standard relativity units

where the Newtonian gravitational constant G is set equal

to 1. The Weyl tensor Cmnlr is just the Riemann tensor with

factors of the Ricci tensor subtracted so that it describes

those aspects of the curvature which are not directly related

to the energy-momentum of the matter.

This gives [7]:

k2 � 8pm2

ð2bþ 4cÞTmlkmkl þ 8c

m2Cmnlrk

mklanar ¼ 0; ð38Þ

which splits the light-cone corrections into two pieces, the

first representing the effect of matter and the second the

gravitational field. The matter contribution turns out to be

relatively universal [7 – 11]; the same formula accommo-

dates the effects on photon propagation due to background

magnetic fields, Casimir cavities, finite temperature envir-

onments, etc. It involves a projection of the energy-

momentum tensor whose sign is fixed by the weak-energy

condition, and in all non-gravitational situations this

correction leads to a timelike p2, i.e. a reduction in the

speed of light. The second contribution, however, is unique

to gravity.

It is immediately clear that the contribution of the Weyl

tensor to the light-cone shift depends on the photon

polarisation am. This interaction therefore produces a

polarisation-dependent shift in the velocity of light: gravita-

tional birefringence.Moreover, it hasa special propertywhich

is not immediately evident but can be proved simply using

symmetry properties of the Weyl tensor (see [7]). The two

physical transverse polarisations give a contribution of the

same magnitude but of opposite sign. This has a striking

consequence: if we consider spacetimes (such as the exterior

region of black holes) with vanishing Ricci tensor (since

Tmn=0), so that theWeyl interaction is the only contribution

to the light-cone shift, it follows that if one polarisation

produces a timelike shift in p2, the other must necessarily

produce a spacelike shift. That is, the light cone (38)

necessarily predicts superluminal velocities!

5.2. Cosmological and black hole spacetimes

To understand these new phenomena in more detail, we

now look at two specific examples. First, we consider the

Friedmann-Robertson-Walker (FRW) spacetime which

describes standard big-bang cosmology. This has a vanish-

ing Weyl tensor so only the energy-momentum term in

equation (38) contributes to the velocity shift. Second, we

consider a Ricci flat case, where only the Weyl term

contributes, viz. the Schwarzschild spacetime describing a

non-rotating black hole.

The FRW metric is

ds2 ¼ dt2 � R2ðtÞh dr2

1� kr2þ r2ðdy2 þ sin2ydf2Þ

i; ð39Þ

where R(t) is the scale factor and k=0, + 1 determines

whether the 3-space is flat, open or closed. The energy-

momentum tensor is

Tmn ¼ ðrþ PÞnmnn � Pgmn; ð40Þ

with nm specifying the time direction in a comoving

orthonormal frame. r is the energy density and P is the

pressure, which in a radiation-dominated era are related by

r – 3P=0.

509Quantum gravitational optics

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 9: Quantum gravitational optics

Since the FRW metric is Weyl flat, the light cone (38) is

simply

k2 ¼ zTmnkmkn; ð41Þ

where z ¼ 8pm2 ð2bþ 4cÞ. The corresponding phase velocity

uph ¼ 1þ 1

2zðrþ PÞ; ð42Þ

is independent of polarisation and is found to be super-

luminal [1].

At first sight, this may look surprising given that k24 0,

its sign fixed by the weak energy condition Tmnkmkn5 0.

However, if instead we consider the momentum along the

rays, p m ¼ Gmnkn, we find

p2 ¼ gmnpmpn ¼ �zðrþ PÞðp0Þ2; ð43Þ

and

uray ¼ jpjp0

¼ 1þ 1

2zðrþ PÞ: ð44Þ

The effective metric G ¼ G�1 is (in the orthonormal frame)

G ¼1þ zr 0 0 0

0 �ð1� zPÞ 0 00 0 �ð1� zPÞ 00 0 0 �ð1� zPÞ

0BB@

1CCA:

ð45Þ

In this case, therefore, we find equal and superluminal

velocities uph= uray and p25 0 is manifestly spacelike as

required.

For a second example [1, 12, 13], consider the Ricci-flat

Schwarzschild spacetime, with metric

ds2 ¼�1� 2M

r

�dt2 �

�1� 2M

r

��1

dr2 � r2ðdy2 þ sin2ydf2Þ:ð46Þ

In the classical theory, with k2=0, there is a special

solution of the null geodesic equations describing a light

ray in a circular orbit with specified radius r=3M. Using

the new light cone (38) and choosing km and am to represent

photons in a circular orbit with transverse polarisations, we

find

k2 ¼ 8c

m2

3M

r3ðk0Þ2; ð47Þ

where k0 is the time component of the wave vector

in an orthonormal frame. The phase velocity is

therefore

uph ¼ k0

k3¼ 1 4c

m2

3M

r3; ð48Þ

and evaluating at r=3M gives the phase velocity of the

new circular orbit, up to O(a). Crucially, we see that one ofthe polarisations has a superluminal velocity.

The radius of the circular photon orbit is changed to

r=3M+O(a), and similarly the location of the effective

ergosphere for the rotating black hole (Kerr metric) is

shifted by polarisation-dependent O(a) corrections due to

the modified speed of light [13]. However, the event horizon

itself has a special status [7, 14]. It can be shown that for a

normal-directed null vector km, both terms in the modified

light cone (38) vanish [15]. It seems that while in general the

quantum-corrected light cones differ from the geometric

null cones, the geometric event horizon remains a true

horizon for the propagation of physical photons.

5.3. Observability?

A similar calculation in the Schwarzschild metric shows

that the angle of deflection of light around a star (one of the

classic tests of general relativity) is modified. If the classical

deflection angle is Dj, then the quantum shift is given by [1]

dðDfÞ ¼ 8c

m2

1

d2f; ð49Þ

where d is the distance of closest approach. This gravita-

tional birefringence effect would also mean that there is in

principle a polarisation dependence in gravitational lensing.

Is this observable?

The problem is that the magnitude of all these quantum

gravitational optics phenomena is incredibly small for

astrophysical scales. Recall that the order of magnitude of

the shift in the speed of light is Oðal2c=L2Þ for a typical

curvature scale L. It is therefore suppressed by the square

of the ratio of a quantum scale (lc) to an astrophysical

curvature scale L. This is tiny 3 – for a solar mass black

hole the correction (48) to the speed of light is

Oðam4pl=ðm2M2ÞÞ � Oð10�34Þ, while the correction (49) to

the bending of light for solar parameters is only of

O(10– 47). Of course this is not surprising – it follows

immediately from the discussion of the physics of vacuum

polarisation in section 3 – and is typical of all phenomena

in the theory of quantum fields in curved spacetime, such as

the famous example of Hawking radiation from black

3To see this, we restore the Newton constant into the metric (46) so thatM?GM. At the critical circular orbit r=3GM, the velocity shift (48) istherefore of O(a/m2G2M2). However, this is still in units where �h= c=1,so that G ¼ 1=m2

pl. Recall that with all fundamental constants manifest, thePlanck mass is given by mpl ¼

ffiffiffiffiffiffiffiffiffiffiffi�hc=G

p. The dimensions are then clearly

consistent and the stated result follows.

510 G. M. Shore

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 10: Quantum gravitational optics

holes. The effect only becomes large for quantum-scale

black holes, which may have arisen in the very early

universe. Like Hawking radiation, therefore, the main

interest in this phenomenon of quantum-induced super-

luminal propagation is probably conceptual: the existence

of thermal Hawking radiation forces us to reconsider the

consistency of quantum mechanics with gravity, while this

superluminal effect challenges our understanding of time

and causality.

Similarly, in the case of the FRW spacetime, the change

in the speed of light only becomes appreciable when the

curvature becomes of order the quantum scale, at which

point our weak-field approximation breaks down and we

can only guess at the possible effects. Nevertheless, in the

radiation dominated era, where rðtÞ ¼ 332p t

�2, we have

uph ¼ 1þ 1

16pzt�2 ð50Þ

which, as already observed in reference [1], increases

towards the early universe. Although this expression is

only reliable in the perturbative weak-field regime, it is

nonetheless intriguing that QED predicts a rise in the speed

of light in the early universe. It is interesting to speculate

whether this superluminal effect persists for high curvatures

near the initial singularity and whether it could play a role

in resolving the horizon problem in cosmology.4

Beyond these potential cosmological or astrophysical

effects, there is a potential obstruction in principle to the

observation of these superluminal phenomena. The argu-

ment runs as follows [1]. If we assume that a typical time

over which propagation can be followed is characterised by

the curvature scale L, then the length difference between

paths corresponding to different polarisations is al2c=L.Very loosely, let us say that observability requires this to be

O(l), where l is the wavelength of the light. The condition

for observability is therefore al2c=ðlLÞ > 1. But the deriva-

tion required not only the weak-field assumption L lcbut also assumed (because of the truncated derivative

expansion in the action) that the photon frequency was

small, i.e. l lc. Now in practice, spectroscopic techni-

ques allow very much more sensitive detection so the O(1)

on the right hand side of the observability criterion could

be many orders of magnitude less (indeed factors of 1010

are achieved in the interferometry used in gravitational

wave detectors). Nevertheless, this does mean that to ensure

that the superluminal phenomenon is genuinely observable

as a matter of principle and that there is no absolute

obstruction to its measurement, we need to extend the

derivation of the light cone condition beyond the low-

frequency approximation. This is addressed in section 7,

where we study the whole question of dispersion.

6. Causality

The ability to send signals ‘faster than light’ is inextricably

linked in most people’s minds with the possibility of time

travel. However, this identification is founded mainly on

intuition arising from special relativity. As we shall see, the

correspondence between superluminal propagation and

causality in general relativity is rather more subtle.

6.1. Superluminal propagation in special and general

relativity

We therefore begin by considering superluminal propaga-

tion in special relativity. The first important observation is

that given a superluminal signal we can always find a

reference frame in which it is travelling backwards in time.

This is illustrated in figure 2.

Suppose we send a signal from O to A at speed u4 1 (in

c=1 units) in frame S with coordinates (t, x). In a frame S0

moving with respect to S with velocity u > 1u, the signal

travels backwards in t’ time, as follows immediately from

the Lorentz transformation. To see this in detail, recall that

from the Lorentz transformations we have t0A ¼ gðuÞtAð1� uuÞ and x 0

A ¼ gðuÞxAð1� uu Þ. For the situation realised

in figure 2, we require both x�AA > 0 and t�AA < 0, that is1u < u < u, which admits a solution only if u4 1.

The important point to be considered is that this by

itself does not necessarily imply a violation of causality.

For this, we require that the signal can be returned from

A to a point in the past light cone of O. However, if we

return the signal from A to B with the same speed in

frame S, then of course it arrives at B in the future cone

of O. The situation is physically equivalent in the Lorentz

boosted frame S 0 – the return signal travels forward in t’time and arrives at B in the future cone of O. This, unlike

the assignment of spacetime coordinates, is a frame-

independent statement.

The problem with causality arises from the scenario

illustrated in figure 3. Clearly, if a backwards-in-time signal

OC is possible in frame S, then a return signal sent with the

same speed will arrive at D in the past light cone of O

creating a closed time loop OCDO. The crucial point is that

local Lorentz invariance of the laws of motion implies that

if a superluminal signal such as OA is possible, then so is

one of type OC5, since it is just given by an appropriate

4Recall that the ‘horizon problem’ is the result that our presently-observeduniverse is extremely homogeneous and isotropic even though it comprisesregions which according to standard big-bang cosmology could never havebeen in causal contact. The conventional explanation is that at very earlytimes, the universe underwent a period of exponentially rapid growth,called inflation, so that our entire observable universe evolved from just onecausally connected region. An alternative, first suggested to me by I.Drummond [private communication] and subsequently exploited in VSLtheories (see section 8), is that the speed of light itself was much greater inthe early universe, thereby enabling light contact between regions we wouldnormally think of as causally disconnected.

511Quantum gravitational optics

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 11: Quantum gravitational optics

Lorentz boost. The existence of global inertial frames then

guarantees the existence of the return signal CD (in

contrast to the situation in figure 2 viewed in the S 0 frame).

The moral is that both conditions must be met in order

to guarantee the occurrence of unacceptable closed time

loops – the existence of a superluminal signal and global

Lorentz invariance. Of course, since global Lorentz

invariance (the existence of global inertial frames) is the

essential part of the structure of special relativity, we

recover the conventional wisdom that in this theory,

superluminal propagation is indeed in conflict with

causality.

So much for special relativity. The situation is,

however, crucially different in general relativity. This is

formulated on the basis of the weak equivalence principle,

which we understand here as the statement that a local

inertial frame exists at each point in spacetime. This

implies that spacetime in general relativity is a Rieman-

nian manifold. However, local Lorentz invariance alone is

not sufficient to establish the link between superluminal

propagation and causality violation. This is usually

established by adding a second, dynamical, assumption.

The strong equivalence principle (SEP) states that the laws

of physics should be identical in the local frames at

different points in spacetime, and that they should reduce

to their special relativistic forms at the origin of each local

frame. It is the SEP which takes over the role of the

existence of global inertial frames in special relativity in

establishing the incompatibility of superluminal propaga-

tion and causality.

O

A

B

t

x O

A

t’

x’

B

Figure 2. A superluminal (u4 1) signal OA which is forwards in time in frame S is backwards in time in a frame S 0 moving relative to

S with speed u > 1u. However, the return path with the same speed in S arrives at B in the future light cone of O, independent of

the frame.

t

xO

C

D

Figure 3. A superluminal (u4 1) signal OC which is backwards

in time in frame S is returned at the same speed to point D in

the past light cone of O, creating a closed time loop.

5For example, suppose we have a conventional tachyon with a spacelikemomentum p2=–m25 0. Setting pm= dxm/ds so that its velocity isutach= djxj/dt= jpj/p0, we find utach ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffi1þ m2

E2

qwhere we have used the

notation E= p0 and as usual set c=1. The velocity therefore depends onthe energy (note that the dispersion relation p2+m2=0, while Lorentzinvariant, is not homogeneous in p), so by an appropriate Lorentztransformation we can change the superluminal tachyon velocity at will.The general result is that if the dynamical equations are invariant underlocal Lorentz transformations and superluminal motion of type OA (figure2) is possible, then so is ‘backwards-in-time’ motion of type OC (figure 3).

512 G. M. Shore

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 12: Quantum gravitational optics

However, unlike the weak equivalence principle, which

underpins the essential structure of general relativity, the

SEP is merely a simplifying assumption about the

dynamics of matter coupled to gravitational fields.

Mathematically, it requires that matter or electromagnet-

ism is minimally coupled to gravity, i.e. with interactions

depending only on the Christoffel symbols but not the

local curvature. This ensures that at the origin of a local

frame, where the Christoffel symbols may be Lorentz

transformed locally to zero, the dynamical equations

recover their special relativistic form. In particular, the

SEP is violated by interactions which explicitly involve the

curvature, such as those occurring in the quantum

Drummond-Hathrell action and the consequent modified

light cones.

We discuss below the question of whether this specific

realisation of superluminal propagation is in conflict

with causality, using the concept of stable causality

described in reference [16]. Notice though that by

violating the SEP, we have evaded the necessary

association of superluminal motion with causality viola-

tion that held in special relativity. Referring back to the

figures, what is established is the existence of a signal of

type OA, which as we saw, does not by itself imply

problems with causality even though frames such as S 0

exist locally with respect to which motion is backwards

in time. However, since the SEP is broken, even if a

local frame exists in which the signal looks like OC, it

does not follow that a return path CD is allowed. The

signal propagation is fixed, determined locally by the

spacetime curvature.

6.2. Stable causality

One special case where causality is realised in a

particularly simple way in general relativity is where the

whole spacetime can be split up into a set of spacelike

slices. As time evolves, a particle moves steadily along a

timelike path from one slice to the next. Clearly, this

excludes the possibility of looping back to the beginning,

i.e. no closed timelike curves exist. Technically, such

spacetimes are called ‘globally hyperbolic’. It is not hard

to imagine that the same basic structure could be

preserved if we use the effective metric Gmn to define

‘spacelike’ and ‘timelike’, rather than the true spacetime

metric gmn, especially if Gmn only differs from gmn by small

terms of O(a). However, even for small differences, this is

a potentially subtle global question and the preservation

of the globally hyperbolic structure cannot be guaranteed

a priori.

We therefore need some advanced theorems of general

relativity to attack this problem in general. In fact, the

clearest general criterion for causality involves the

concept of stable causality discussed, for example, in

the monograph of Hawking and Ellis [16] (see also [17]).

Proposition 6.4.9 states the required definition and

theorem:

. A spacetime manifold (M, gmn) is stably causal if

the metric gmn has an open neighbourhood such that

M has no closed timelike or null curves with re-

spect to any metric belonging to that neighbour-

hood.

. Stable causality holds everywhere on M if and only

if there is a globally defined function f whose gradi-

ent Dmf is everywhere non-zero and timelike with re-

spect to gmn.

The proof of this involves analysis beyond the scope of this

review. The meaning of the theorem, however, is quite

straightforward. It simply asserts that the absence of

causality violation in the form of closed timelike or

lightlike curves is assured if we can find a globally defined

function f whose gradient is timelike with respect to the

effective metric Gmn for light propagation. f then acts as a

global time coordinate.

We can immediately give one example where this

condition does hold. This is the FRW spacetime

discussed in section 4.2. In this case, the effective metric

(in the unperturbed orthonormal frame) is given by (45).

We simply use the cosmological time coordinate t as the

globally defined function f. We need only then check

that its gradient Dmt defines a timelike vector with

respect to the effective metric Gmn. This is true provided

G004 0, which is certainly satisfied by equation (45). So

at least in this case, superluminal propagation is

compatible with causality.

7. Gravity’s rainbow

So far, our discussion of photon propagation has been

based on the low-frequency approximation arising from

the effective action (27). However, we know that this

cannot be the whole story. In this section, we investigate

dispersion – how the speed of light depends on its

frequency – and explain carefully why the high-frequency

limit is critical for causality. We begin with a review of

some basic properties of light propagation in classical

optics.

7.1. The ‘speeds of light’

An illuminating discussion of wave propagation in a

simple dispersive medium is given in the classic

work by Brillouin [18]. This considers propagation

of a sharp-fronted pulse of waves in a medium

with a single absorption band, with refractive index

n(o):

513Quantum gravitational optics

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 13: Quantum gravitational optics

n2ðoÞ ¼ 1� a2

o2 � o20 þ 2ior

; ð51Þ

where a and r are constants and o0 is the characteristic

frequency of the medium. Five6 distinct velocities are

identified: the phase velocity uph ¼ ojkj ¼ <� 1

nðoÞ�, group

velocity ugp ¼ dodjkj, signal velocity usig, energy-transfer velo-

city uen and wavefront velocity uwf, with precise definitions

related to the behaviour of contours and saddle points in

the relevant Fourier integrals in the complex o-plane. Theirfrequency dependence is illustrated in figure 4. Notice that

here and in what follows, we use < and = to denote real and

imaginary parts respectively.

As the pulse propagates, the first disturbances to arrive

are very small amplitude waves, ‘frontrunners’, which

define the wavefront velocity uwf. These are followed

continuously by waves with amplitudes comparable to the

initial pulse; the arrival of this part of the complete

waveform is identified in reference [18] as the signal velocity

usig. As can be seen from figure 4, it essentially coincides

with the more familiar group velocity for frequencies far

from o0, but gives a much more intuitively reasonable sense

of the propagation of a signal than the group velocity,

whose behaviour in the vicinity of an absorption band is

relatively eccentric.

In passing, notice that it is the group velocity which is

measured in quantum optics experiments which find light

speeds of essentially zero [19] or many times c [20]. For

example, in the experiments of Vestergaard Hau and

colleagues at Harvard, the group velocity of a light pulse

is reduced almost to zero by shining tuned lasers on a

cloud of ultra-cold sodium atoms. The set-up is designed

to produce an effective refractive index of the cloud of

the form shown in figure 5. A very clear explanation of

how this is achieved by manipulating the states of the

sodium atoms in the gas cloud and exploiting quantum

superposition is given in reference [19]. Now, some

simple algebra shows that the group velocity can be

expressed in terms of the slope of the refractive index:

ugp ¼ �nþ o dndo

��1. The group velocity therefore becomes

small in a region where the slope of the refractive index

is large. In the ‘slow-light’ experiments, the central

section of the curve n(o) in figure 5 is made as steep

as possible and the frequency of the probe laser tuned to

this frequency range, forcing the group velocity towards

zero.

Returning to figure 4, it is clear that the phase velocity

itself also does not represent a ‘speed of light’ relevant for

considerations of signal propagation or causality. The

appropriate velocity to define light cones and causality is in

fact the wavefront velocity uwf. (Notice that in figure 4, uwfis a constant, equal to c, independent of the frequency or

details of the absorption band). This is determined by the

boundary between the regions of zero and non-zero

disturbance (more generally, a discontinuity in the first or

higher derivative of the field) as the pulse propagates.

Mathematically, this definition of wavefront is identified

with the characteristics of the partial differential equation

governing the wave propagation [21]. Our problem is

therefore to determine the velocity associated with the

characteristics of the wave operator derived from the

(velocity)1

ww00

c

signal

phase

group

1

Figure 4. Sketch of the behaviour of the phase, group and signal

velocities with frequency in the model described by the refractive

index equation (51). The energy-transfer velocity (not shown) is

always less than c and becomes small near o0. The wavefront

speed is identically equal to c.

1

0

w

n(w)

Figure 5. The effective refractive index of the cloud of ultra-cold

sodium atoms used in the ‘slow light’ experiments.

6In fact, if we take into account the distinction discussed in section 4between the phase velocity uph and the ray velocity uray, and include thefundamental speed of light constant c from the Lorentz transformations,we arrive at seven distinct definitions of ‘speed of light’.

514 G. M. Shore

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 14: Quantum gravitational optics

modified Maxwell’s equations of motion appropriate to the

new effective action.

Notice that a very complete and rigorous discussion of

the wave equation in curved spacetime has already been

given by Friedlander in reference [22], where it is proved

(Theorem 3.2.1) that the characteristics are simply the null

hypersurfaces of the spacetime manifold, in other words

that the wavefront always propagates with the fundamental

speed c. However, this discussion assumes the standard

form of the (gauge-fixed) Maxwell’s wave equation and

does not cover the modified wave equation (29), precisely

because of the extra curvature couplings which lead to the

effective metric Gmn and superluminal propagation.

Instead, the key result which allows a derivation of the

wavefront velocity is derived by Leontovich [23]. In this

paper, an elegant proof is presented for a very general set of

PDEs that the wavefront velocity associated with the

characteristics is identical to the o ! 1 limit of the phase

velocity, i.e.

uwf ¼ limo!1

ojkj ¼ lim

o!1 uphðoÞ: ð52Þ

The proof is rather formal (see [5] for further details) but is

of sufficient generality to apply to our discussion of photon

propagation using the modified effective action of section 4.

Notice in particular that this implies uwf is indeed a

constant, independent of the frequency.

The wavefront velocity in a gravitational background is

therefore not given a priori by c. Taking vacuum polarisa-

tion into account, there is no simple non-dispersive medium

corresponding to the vacuum of classical Maxwell’s theory

in which the phase velocity represents a true speed of

propagation; for QED in curved spacetime, even the

vacuum is dispersive. In order to discuss causality, we

therefore have to extend the original Drummond-Hathrell

results for uph(o*0) to the high frequency limit

uphðo ! 1Þ, as already emphasised in their original work.

A further subtle question arises if we write the standard

optics dispersion relation for the refractive index n(o) in the

limit o ! 1:

nð1Þ ¼ nð0Þ � 2

p

Z 1

0

doo

= nðoÞ½ �: ð53Þ

For a conventional dispersive medium, =½nðoÞ� > 0, which

implies that nð1Þ < nð0Þ, or equivalently uphð1Þ > uphð0Þ.Evidently this is satisfied by figure 4. The key question

though is whether the usual assumption of positivity of

=½nðoÞ� holds in the present situation of the QED vacuum

in a gravitational field. If so, then (as already noted in

reference [1]) the superluminal Drummond-Hathrell results

for uph(0) would actually be lower bounds on the all-

important wavefront velocity uphð1Þ. However, it is not

clear that positivity of =½nðoÞ� holds in the gravitational

context. Indeed it has been explicitly criticised by Dolgov

and Khriplovich in references [24, 25], who point out that

since gravity is an inhomogeneous medium in which beam

focusing as well as diverging can happen, a growth in

amplitude corresponding to =½nðoÞ� < 0 is possible. The

possibility of uphð1Þ < uphð0Þ, and in particular

uphð1Þ ¼ c, therefore does not seem to be convincingly

ruled out by the dispersion relation equation (53).

7.2. Light in magnetic fields

Before confronting the gravitational case, we can

build up some intuition on dispersion by looking at

a closely related situation – the propagation of light

through a background magnetic field. In this case, the

analogue of the low-frequency, weak-field Drummond-

Hathrell action for gravity is the Euler-Heisenberg

action:

G ¼Z

dx

� 1

4F mnF

mn þ 1

m4

pF mnF lrF

mlF nrþ

q�F mnF

mn�2�

;

ð54Þ

where p ¼ 790 a

2 and q ¼ � 136 a

2. For a constant magnetic

field, a familiar geometric optics analysis produces the

following results for the phase velocity of photons moving

transverse to the background field:

k2 þ 8

m2

hð2pþ 4qÞk; p?

iTmnk

mkn ¼ 0 ð55Þ

where Tmu ¼ F ml F lu � 1

4 gmnF lrFlr is the electromagnetic

energy-momentum tensor and the suffices k;? indicate the

two polarisations. This should be compared with equation

(38) for the gravitational case. Notice that for electro-

magnetism, birefringence appears already in the energy-

momentum tensor term, essentially because the lowest

order terms in the Euler-Heisenberg action contributing to

(55) are of 4th order in the background fields, i.e. O(F4),

compared to the 3rd order terms of O(RFF) in the

Drummond-Hathrell action. Substituting for p,q we find

the well-known results for the low-frequency limit of the

phase velocity:

uphðo ! 0Þ � 1� a4p

�eBm

�2h1445k

;8

45?

i: ð56Þ

As expected, uph5 1 for both polarisations.

The refractive index for photons of arbitrary frequency

has been calculated explicitly by Tsai and Erber [33, 34].

They derive an effective action

515Quantum gravitational optics

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 15: Quantum gravitational optics

G ¼ �Z

dx

�1

4F mnF

mn þ 1

2AmMmnA

n�; ð57Þ

where Mmn is a differential operator acting on the

electromagnetic field An. Writing the corresponding equa-

tion of motion and making the geometric optics ansatz as

usual, we find the light cone condition

k2 �MmnðkÞaman ¼ 0: ð58Þ

Denoting Mmnaman by Mk; M? for the two polarisations,

the complete expression for the birefringent refractive index

is:

uphk;?ðoÞ ¼ 1þ 1

2o2Mk;?

¼ 1þ a4p

�eBm2

�2 Z 1

�1

du

Z 1

0

ds sNk;?ðu; zÞe�is�1þs2O2Pðu;zÞ

�;ð59Þ

where z ¼ eBm2 s and O ¼ eB

m2om. The functions N and P are

known exactly.

In the weak field, low frequency limit, we can disregard

the function P (since O is small) and consider only the

lowest term in the expansion of N in powers of z. This

reproduces the results for uphðo ! 0Þ shown in equation

(56). The weak field, high frequency limit is analysed in

reference [34]. It is shown that

uphk;?ðo ! 1Þ � 1þ a4p

�eBm

�2�ck; c?

�O�4

3; ð60Þ

where the numerical constants ½ck; c?� are known.

The complete frequency dependence of uph(o), or

equivalently n(o), is therefore known and shows exactly

the features found in the simple absorption model described

above. In particular, the phase velocity uph(o) begins less

than 1 at low frequencies, showing birefringence but

conventional subluminal behaviour. In the high frequency

limit, however, the phase velocity approaches c from the

superluminal side with a o�43 behaviour.

The origin of this non-analytic high-frequency behaviour

is of particular interest. It arises from the phase factor

exp½�is3O2Pðu; zÞ� in the effective action. The absorption

resonance region in the graph of the refractive index n(o)occurs for frequencies o such that O*1. For high

frequencies, even at weak fields, we need O ! 1. Clearly,

these aspects of the behaviour of n(o) are invisible to a

simple expansion of the effective action in low powers of

the field strength B, even keeping terms of all orders in

derivatives. The conclusion is that a complete analysis of

resonant and high frequency propagation requires the

evaluation of the non-perturbative (in a field strength

expansion) phase factor in equation (59). As we now see,

this poses severe problems for the analysis of high

frequency propagation in the case of a background

gravitational field.

7.3. Dispersion and gravity

Setting aside this concern for the moment, we now return to

gravity and begin our study of dispersion by constructing

the generalisation of the Drummond-Hathrell effective

action containing all orders in derivatives, while retaining

only terms of O(RFF) in the curvature and field strength.

This action was recently derived in reference [4]. The result

is:

G ¼Z

dxffiffiffiffiffiffiffi�g

p � 1

4FmnF

mn þ 1

m2

�DmF

ml G0�!

DnFnl

þ G1�!

RFmnFmn þ G2

�!RmnF

mlFnl þ G3

�!RmnlrF

mnFlr�

þ 1

m4

�G4�!

RDmFmlDnF

nl þ G9

�!RmnlrDsF

srDlFmn

þ G5�!

RmnDlFlmDrF

rn þ G6�!

RmnDmFlrDnFlr

þ G7�!

RmnDmDnFlrFlr þ G8

�!RmnD

mDlFlrFrn�

: ð61Þ

This was found by adapting a background field action valid

to third order in generalised curvatures due to Barvinsky et

al. [26] (see also reference [27]) and involves re-expressing

their more general result in manifestly local form by an

appropriate choice of basis operators.

In this formula, the G!

nðn51Þ are form factor functions

of three operators:

Gn�! � Gn

�D2ð1Þ

m2;D2

ð2Þm2

;D2

ð3Þm2

�; ð62Þ

where the first entry ðD2ð1ÞÞ acts on the first following

term (the curvature), etc. G!

0 is similarly defined as a

single variable function. These form factors are found

using heat kernel methods and are given by ‘proper time’

integrals of known algebraic functions. Their explicit

expressions can be found in reference [4]. Evidently,

equation (61) reduces to the Drummond-Hathrell action

if we neglect all the higher order derivative terms.

The next step is to derive the equation of motion

analogous to equation (29) from this generalised effective

action and to apply geometric optics to find the corre-

sponding light cone. This requires a very careful analysis of

the relative orders of magnitudes of the various contribu-

tions to the equation of motion arising when the factors of

D2 in the form factors act on the terms of O(RF). These

subtleties are explained in detail in reference [3]. The final

result for the new effective light cone has the form

516 G. M. Shore

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 16: Quantum gravitational optics

k2 � 8pm2

F�k:Dm2

�Tmlk

mkl þ 1

m2G�k:Dm2

�Cmnlrk

mklanar ¼ 0;

ð63Þ

where F and G are known functions with well-understood

asymptotic properties [3]. Clearly, for agreement with

equation (27), we have F(0)=2b+4c, G(0)=8c.

The novel feature of this new light cone condition is that

F and G are functions of the operator k.D acting on the

Ricci and Riemann tensors.7 So although the asymptotic

behaviour of F and G as functions is known, this

information is not really useful unless the relevant

curvatures are eigenvalues of the operator. On the positive

side, however, k.D does have a clear geometrical inter-

pretation – it simply describes the variation along a null

geodesic with tangent vector km.

The utility of this light cone condition therefore seems to

hinge on what we know about the variations along null

geodesics of the Ricci and Riemann (or Weyl) tensors.

Unfortunately, we have been unable to find any results in

the general relativity literature which are valid in a general

spacetime. To try to build some intuition, we have therefore

again looked at particular cases. The most interesting

example in this case is photon propagation in the Bondi-

Sachs metric [29, 30] which we recently studied in detail

[31].

The Bondi-Sachs metric describes the gravitational

radiation from an isolated source. The metric is

ds2 ¼ �Wdu2 � 2e2bdudrþ r2hijðdxi �UiduÞðdxj �UjduÞ;ð64Þ

where

hijdxidxj ¼ 1

2ðe2g þ e2dÞdy2 þ 2 sinhðg� dÞsinydydf

þ 1

2ðe�2g þ e�2dÞsin2ydf2:

ð65Þ

The metric is valid far from the source, i.e. for large r. The

family of hypersurfaces u= const are null, i.e. gmn@mu@nu=0. Their normal vector ‘m satisfies

‘m ¼ @mu ) ‘2 ¼ 0; ‘mDm‘n ¼ 0: ð66Þ

The curves with tangent vector ‘m are therefore null

geodesics; the coordinate r is a radial parameter along

these rays and is identified as the luminosity distance. The

six independent functions W, b, g, d and Ui characterising

the metric have expansions in 1r in the asymptotic region,

the coefficients of which describe the various features of the

gravitational radiation.

In the low frequency limit, the light cone is given simply

by equation (38) with Tmn=0. The velocity shift is quite

different for the case of outgoing and incoming photons

[31]. For outgoing photons, km ¼ ol m, and the light cone,

written in Newman-Penrose notation8 is

k2 ¼ 4co2

m2

�C0 þC�

0

�� O

�1r5

�; ð67Þ

while for incoming photons, km=onm,

k2 ¼ 4co2

m2

�C4 þC�

4

�� O

�1r

�: ð68Þ

Now, it is a special feature of the Bondi-Sachs spacetime

that the absolute derivatives of each of the Weyl scalars C0,

. . ., C4 along the ray direction ‘m vanish, i.e. C0, . . ., C4 are

parallel transported along the rays [30, 32]. In this case,

therefore, we have:

‘ �DC0 ¼ 0; ‘ �DC4 ¼ 0; ð69Þ

but there is no equivalent simple result for either n�D C4 or

n�D C0.

Although it is just a special case, equation (69) never-

theless leads to a remarkable conclusion. The full light cone

condition equation (63) applied to outgoing photons in the

Bondi-Sachs spacetime now reduces to

k2 o2

2m2G�0��

C0 þC�0

�¼ 0; ð70Þ

since ‘ �DC0 ¼ 0. In other words, the low-frequency

Drummond-Hathrell prediction of a superluminal phase

velocity uph(0) is exact for all frequencies. There is no

dispersion, and the wavefront velocity uphð1Þ is indeed

superluminal.

This is potentially a very important result. It appears to

show that there is at least one example in which the

wavefront truly propagates with superluminal velocity. If

so, quantum effects would indeed have shifted the light

cone into the geometrically spacelike region.

However, we now have to confront the problem raised

above in the discussion of light propagation in background

7Note that because these corrections are already of O(a), we can freely usethe usual Maxwell’s relations k.Dkn=0 and k.Dan=0 in these terms; weneed only consider the effect of the operator k.D acting on Rmn and Rmnlr.

8Details of the Newman-Penrose formalism can be found, e.g., in references[7,31]. The essential idea is to construct a basis of null vectors ‘m, nm, mm and�mmm, and refer the components of the Ricci and Weyl tensors to this basis.The 10 independent components of the Weyl tensor can then be written as aset of 5 complex scalars, denoted C0, . . ., C4. For example,C0 ¼ �Cmnlr‘

mmn‘lmr. In the example above, we identify the momentumdirection for outgoing photons with ‘m; mm is then the complex linearcombination ak þ ia?of the physical transverse polarisations.

517Quantum gravitational optics

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 17: Quantum gravitational optics

magnetic fields. If the gravitational case is similar, this

would imply that the modified light cone can be written

heuristically as

k2 þ ap

Z 1

0

ds Nðs;RÞe�is�1þs2O2Pðs;RÞ

�¼ 0; ð71Þ

where both N and P can be expanded in powers of

curvature, and derivatives of curvature, presumably asso-

ciated with factors of o as in the last section. The frequency

dependent factor O would be O � Rm2

om � O

�l3

lL2

�, where ‘R’

denotes some generic curvature component and L is the

typical curvature scale. If this is true, then an expansion of

the effective action to O(RFF), even including higher

derivatives, would not be sufficient to reproduce the full,

non-perturbative contribution expð�is3O2PÞ. The Drum-

mond-Hathrell action would correspond to the leading

order term in the expansion of equation (71) in powers of Rm2

neglecting derivatives, while our improved effective action

(61) sums up all orders in derivatives while retaining the

restriction to leading order in curvature.

The omission of the non-perturbative contribution

would be justified only in the limit of small O, i.e. forl3clL2 � 1. Neglecting this therefore prevents us from acces-

sing the genuinely high frequency limit l ! 0 needed to

find the asymptotic limit uphð1Þ of the phase velocity.

Moreover, assuming equation (71) is indeed on the right

lines, it also seems inevitable that for high frequencies (large

O) the rapid phase variation in the exponent will drive the

entire heat kernel integral to zero, ensuring the wavefront

velocity uwf= c.

At present, it is not clear how to make further progress.

The quantum field theoretic calculation required to find

such non-perturbative contributions to the effective action

and confirm an expð�is3O2PÞ structure in equation (71)

appears difficult, although some technical progress in this

area has been made recently in reference [36] and work in

progress [37]. One of the main difficulties is that since a

superluminal effect requires some anisotropy in the

curvature, it is not sufficient just to consider constant

curvature spacetimes. [Recall that the Ricci scalar term in

the effective action (27) does not contribute to the modified

light cone (30)]. A possible approach to this problem, which

would help to control the plethora of indices associated

with the curvatures, might be to reformulate the calculation

of the effective action directly in the Newman-Penrose

basis. On the other hand, perhaps a less ambitious goal

would be to try to determine just the asymptotic form of

the non-perturbative contribution in the O ! 1 limit.

A final resolution of the dispersion problem for QED in

curved spacetime has therefore still to be found. At present,

the most likely scenario appears to be that high-frequency

dispersion is driven by non-perturbative contributions to

the effective action such that the wavefront velocity remains

precisely c. It would then be interesting to see exactly how

the phase velocity behaves as a function of o. What we

have established is that for ‘sub-resonant’ frequencies the

phase velocity may, depending on the polarisation, take

both the conventional form of figure 4 or the superluminal

mirror image where uph(0)4 1. We have even seen in the

Bondi-Sachs example a special case where uph(o)*const4 1 for relatively low frequencies.

However, even if this picture is correct and the light cone

is eventually driven back to k2=0 in the high frequency

limit, the analysis described here still represents a crucial

extension of the domain of validity of the superluminal

velocity prediction of Drummond and Hathrell. Recall

from section 4.3 that the constraint on the frequency for

which the superluminal effect is in principle observable isl2lL 1. Obviously this was not satisfied by the original

o*0 derivation. However, our extension based on the

generalised effective action (61) does satisfy this constraint.

Combining with the restriction l3

lL2 � 1 in which the neglect

of the P type corrections is justified, we see that there is a

frequency range

lcL

llc

l2cL2

; ð72Þ

below the ‘resonance’ region of figure 4, for which our

expression (63) for the modified light cone is valid and

predicts in principle observable effects.

Since this formula allows superluminal corrections to

both the phase and the signal velocities, we conclude that

superluminal propagation has indeed been established as an

observable phenomenon even if, as seems likely, causality

turns out to respected through the restoration of the

standard light cone k2=0 in the asymptotic high frequency

limit.

8. Speculations

In this paper, we have reviewed some of the qualitatively

new phenomena that occur when quantum theory is

introduced into the already rich field of gravitational

optics. The picture that arises is of curved spacetime as an

optical medium, characterised by a novel refractive index

and displaying polarisation-dependent properties such as

birefringence.

Naturally, since these phenomena arise due to intrinsi-

cally quantum field theoretic processes such as vacuum

polarisation, they are extremely small for the curvatures

typical of present day astrophysics. A typical order of

magnitude for the shift in the speed of light around a black

hole, for example, is O½am4pl=ðm2M2Þ�, where m is the

electron mass setting the scale for vacuum polarisation and

M is the black hole mass. Quantum gravitational optics

518 G. M. Shore

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 18: Quantum gravitational optics

(QGO) is therefore only likely to play an important role in

situations of extreme, quantum-scale, curvature such as

may be present in the very early universe or in the vicinity

of primordial black holes. Its main interest, as in the

analogous case of Hawking radiation, is likely to be

primarily conceptual.

Indeed, serious conceptual issues do arise through the

apparent prediction in certain circumstances of a super-

luminal speed of light. We have explained carefully how

this occurs and discussed the subtle questions of inter-

pretation that arise. Although further work remains to be

done to achieve a full understanding, it appears that despite

an anomalous refractive index featuring a superluminal

low-frequency phase (and signal) velocity, the critical high-

frequency limit which determines the characteristics or

wavefront velocity remains equal to c. Nonetheless, we

described how superluminal velocities could be accommo-

dated in the framework of general relativity while preser-

ving the essential notion of stable causality.

All this work was within the relatively uncontroversial

theoretical formalism of quantum electrodynamics in

curved spacetime. To close, we give a brief survey of other

more speculative ideas in the current literature which

overlap in some way with the ideas presented here.

First, we should note that everything that has been said

about the QED effective action arises equally in the low-

energy effective actions derived from string theory. SEP-

violating interactions of O(RFF) naturally occur there too,

with the main difference that the electron scale m becomes

the Planck mass mpl. Evidently this simply reduces the

relative order of magnitude by another huge factor.

One of the key insights in the theory presented here is

that in QGO the usual light cone relation for the photon

momentum becomes

Gmnpmpn ¼ 0: ð73Þ

That is, the propagation of real photons is governed by a

new metric Gmn, which may in general depend not only on

the curvature and its derivatives but also on the photon

direction, polarisation and energy. The physical light cones

are then distinct from the geometric null cones charac-

terised by the spacetime metric gmn. This realisation of QGO

is therefore intrinsically a bimetric theory.

Inspired by this, it is interesting to speculate on

modifications to conventional general relativity in which a

bimetric structure is imposed a priori. Phenomenologically,

this has the advantage that the new scale characterising the

propagation metric can be a free parameter, with the result

that the type of phenomena discussed here could occur on

macroscopically accessible scales. As a result, there is a

strong motivation to search for unusual velocity or

polarisation dependent phenomena in light signals from

cosmological sources. Bimetric theories have a long history,

with a particularly elegant recent construction due to

Drummond [38], who has proposed an extension of general

relativity in which matter couples to its own vierbein, which

is rotated relative to the geometric vierbein according to an

appropriate sigma model dynamics. By modifying gravita-

tional dynamics on an astrophysical scale, such theories

have the potential to offer an alternative to dark matter

models. Other recent work exploiting bimetric theories to

resolve cosmological puzzles can be found in a series of

papers on scalar-tensor gravity by Clayton and Moffat [39,

40].

These theories modify general relativity, and by incor-

porating a bimetric structure they necessarily involve a

varying speed of light (VSL). VSL theories and their

potential for providing an alternative to inflation as a

resolution of cosmological problems such as the horizon

problem have seen a rapid rise in popularity in recent years.

A small selection of relevant papers is included here as an

introduction to this literature for the interested reader [41,

42, 43, 44, 45, 46].

The QED version of QGO discussed in this paper, and

some of the bimetric models, preserve the essential

principles of general relativity, especially the weak equiva-

lence principle and local Lorentz invariance. We can of

course speculate that local Lorentz invariance itself may be

violated. A phenomenological approach to such theories

has been pioneered by Kostelecky, who has constructed a

general extension to the standard model incorporating

Lorentz and CPT9 violating interactions and investigating

their experimental consequences [47]. Of particular interest

here are the following interactions, which affect the

propagation of light:

G ¼ GM þZ

dx

�� 1

4KmnlrF

mnFlr þ 1

2LmEmnlrAnFlr

�: ð74Þ

Here, K and L are Lorentz-violating couplings, which could

plausibly arise in some more fundamental superstring

theory exhibiting spontaneous symmetry breaking of

Lorentz symmetry (analogous to the way the Higgs

mechanism spontaneously breaks gauge symmetry in the

standard model of particle physics). Evidently, the first of

these interactions is the analogue of the SEP-violating

interaction considered in this paper, with the replacement

of the covariant curvature tensor Rmnlr by the coupling

9CPT is the combination of the three discrete symmetries of chargeconjugation (C), parity (P) and time-reversal (T). All are known to bebroken in particle physics. For example, P invariance is broken in weakinteractions such as b decay and T and CP are broken in K 0 � �KK 0 mixing.However, while the individual discrete symmetries may be broken in aparticular model, the combined symmetry CPT is believed to be exact.Indeed, there is a theorem which ensures that CPT is preserved in any local,Lorentz-invariant quantum field theory. CPT violation therefore requiresone of these assumptions to be broken, i.e. either locality as may perhaps bepossible in string theory, or Lorentz invariance.

519Quantum gravitational optics

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 19: Quantum gravitational optics

Kmnlr. The phenomenology of light propagation, and some

(though not all) of the considerations concerning causality

discussed here therefore apply directly to this class of

Lorentz-violating theory. The second term involving Lm has

a topological nature and violates CPT as well as Lorentz

invariance. It does not of course appear in normal QED,

though it can be induced by anomalies involving spacetime

topology [48]. The Kostelecky Lagrangian has been used as

a phenomenological model against which to put limits on

the magnitude of VSL, polarisation-dependent or Lorentz-

violating effects in light propagation. Since a birefringent

speed of light would cause a rotation of the polarisation of

light as it travels through space, the observation of

polarisation from distant quasars and radio galaxies, and

also recent measurements of the polarisation of the cosmic

microwave background, can be used to place limits on the

phenomenological couplings K and L. Details of this work

can be found in reference [49]. Similar interactions are

induced also in non-commutative gauge theories as a

consequence of the Moyal product, with similar phenom-

enological implications. Here, the analogue of the coupling

Kmnlr is constructed from the non-commutativity parameter

ymn defined from [xm, xn]= iymn.Of course, once we permit models with more or less ad

hoc Lorentz violation, then we are free to speculate at will

as to alternative dispersion relations for the propagation of

light, or indeed neutrinos. Some popular recent ideas

include models with a supposed non-realisation of Lorentz

invariance in the effective low-energy field theory derived

from quantum gravity [50, 51, 52] and ‘doubly special’

relativity [53] where the Lorentz transformations are

modified, possibly using a non-linear realisation of the

symmetry [54], so that the Planck length as well as c is an

invariant. Evidently, there is immense freedom to frame

such models within the constraints of experimental data

though it is far less clear how to incorporate them into

complete theories consistent with our understanding of

unified theories of particle physics and quantum gravity.

On the experimental side, particular interest is attached to

light (or neutrinos) received from gamma-ray bursts, since

the extremely high energies of the photons can enhance the

size of the VSL effect for certain Lorentz-violating

dispersion relations.

This rapid tour of admittedly rather speculative models

is hopefully sufficient to indicate the considerable current

interest in looking for anomalies in the standard theory of

light propagation in special and general relativity. How-

ever, as we have shown in this paper, it is not necessary to

invoke ad hoc violations of the fundamental principles of

quantum field theory, general relativity or string theory in

order to discover a new richness of phenomena in the

propagation of light. Whether its primary interest will lie in

the conceptual issues it raises for the consistency of

quantum field theory with gravity, or in the prediction of

experimental anomalies in the speed of light in cosmology,

the field of quantum gravitational optics is promised a

bright future.

Acknowledgments

I am especially grateful to I. Drummond for many

interesting discussions on superluminal propagation and

numerous other colleagues, including A. Dolgov, G.

Gibbons, H. Gies, V. Khoze, S. King, K. Kunze, M.

Maggiore, J. Magueijo, H. Osborn, W. Perkins, S. Sarkar

and G. Veneziano, for their comments on this work. This

research is supported in part by PPARC grant PP/G/O/

2000/00448.

References

[1] Drummond, I. T., and Hathrell, S., 1980, Phys. Rev. D, 22, 343.

[2] Hawking, S. W., 1975, Commun. Math. Phys., 43, 199.

[3] Shore, G. M., 2002, Nucl. Phys. B, 633, 271.

[4] Shore, G. M., 2002, Nucl. Phys. B, 646, 281.

[5] Shore, G. M., 2002, Superluminal Light. In Time and Matter: an

International Colloquium on the Science of Time, edited by I. Bigi and

M. Faessler (Venice), gr-qc/0302116.

[6] Schneider, P., Ehlers, J., and Falco, E. E., 1992, Gravitational Lenses,

(Berlin: Springer-Verlag).

[7] Shore, G. M., 1996, Nucl. Phys. B, 460, 379.

[8] Scharnhorst, K., 1990, Phys. Lett. B, 236, 354.

[9] Barton, G., 1990, Phys. Lett. B, 237, 559.

[10] Latorre, J. I., Pascual, P., and Tarrach, R., 1995, Nucl. Phys. B, 437,

60.

[11] Dittrich, W., and Gies, H., 1998, Phys. Lett. B, 431, 420; 1998, Phys.

Rev., D, 58, 025004.

[12] Daniels, R. D., and Shore, G. M., 1994, Nucl. Phys. B, 425, 634.

[13] Daniels, R. D., and Shore, G. M., 1996, Phys. Lett. B, 367, 75.

[14] Gibbons, G. W., and Herdeiro, C. A. R., 2001, Phys. Rev. D, 63,

064006.

[15] Hawking, S. W., 1972, The Event Horizon. In Les Houches Lectures

1972, Vol 72 Black Holes, edited by C. De Witt, (Gordon and Breach,

New York).

[16] Hawking, S. W., and Ellis, G. F. R., 1973, The Large Scale Structure

of Spacetime, (Cambridge University Press).

[17] Liberati, S., Sonego, S., and Visser, M., 2002, Annals Phys., 298, 167.

[18] Brillouin, L., 1960, Wave Propagation and Group Velocity, (London:

Academic Press).

[19] Hau, L. V., 2001, Scientific American, 7, 66.

[20] Wang, L. J., Kuzmich, A., and Dogoriu, A., 2000, Nature, 406, 277.

[21] Courant, R., and Hilbert, D., 1962, Methods of Mathematical Physics,

Vol II, (New York: Interscience).

[22] Friedlander, F. G., 1975, The Wave Equation on a Curved Spacetime,

(Cambridge University Press).

[23] Leontovich, M. A., 1972, In L.I. Mandelshtam, Lectures in Optics,

Relativity and Quantum Mechanics, Nauka, Moscow (in Russian).

[24] Dolgov, A. D., and Khriplovich, I. B., 1983, Sov. Phys. JETP, 58, 671.

[25] Khriplovich, I. B., 1995, Phys. Lett. B, 346, 251.

[26] Barvinsky, A. O., Gusev, Y. V., Vilkovisky, G. A., and Zhytnikov, V.

V., 1993, Print-93-0274 (Manitoba).

[27] Barvinsky, A. O., Gusev, Y. V., Vilkovisky, G. A., and Zhytnikov, V.

V., 1994, J. Math. Phys., 35, 3525; 1994, J. Math. Phys., 35, 3543;

1995, Nucl. Phys., B, 439, 561.

520 G. M. Shore

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3

Page 20: Quantum gravitational optics

[28] Chandresekhar, S., 1985, The Mathematical Theory of Black Holes,

(Oxford: Clarendon).

[29] Bondi, H., van der Burg, M. G. J., and Metzner, A. W. K., 1962, Proc.

Roy. Soc. A, 269, 21.

[30] Sachs, R. K., 1962, Proc. Roy. Soc. A, 270, 103.

[31] Shore, G. M., 2001, Nucl. Phys. B, 605, 455.

[32] d’Inverno, R. A., 1992, Introducing Einstein’s Relativity, (Oxford:

Clarendon).

[33] Tsai, W., and Erber, T., 1974, Phys. Rev. D, 10, 492.

[34] Tsai, W., and Erber, T., 1975, Phys. Rev. D, 12, 1132.

[35] Adler, S., 1971, Ann. Phys. (N.Y.), 67, 599.

[36] Barvinsky, A. O., and Mukhanov, V. F., 2002, Phys. Rev. D, 66,

065007.

[37] Gusev, Y. V., private communication.

[38] Drummond, I. T., 2001, Phys. Rev. D, 63, 043503.

[39] Clayton, M. A., and Moffat, J. W., 2001, Phys. Lett. B, 506, 177;

astro-ph/0203164.

[40] Moffat, J. W., 2003, Int. J. Mod. Phys. D, 12, 281.

[41] Moffatt, J. W., 1993, Int. J. Mod. Phys. D, 2, 351.

[42] Albrecht, A., and Magueijo, J., 1999, Phys. Rev. D, 59, 043516.

[43] Barrow, J. D., 1999, Phys. Rev. D, 59, 043515.

[44] Barrow, J. D., and Magueijo, J., 1999, Phys. Lett. B, 447, 246.

[45] Clayton, M. A., and Moffatt, J. W., 1999, Phys. Lett. B, 460, 263;

2000, Phys. Lett. B, 447, 269.

[46] Bassett, B. A., Liberati, S., Molina-Paris, C. and Visser, M., 2000,

Phys. Rev. D, 62, 103518.

[47] Colladay, D., and Kostelecky, V. A., 1998, Phys. Rev. D, 58, 116002.

[48] Klinkhamer, F. R., CPT violation: mechanism and phenomenology.

In Seventh International Wigner Symposium, edited by M. E. Noz, hep-

th/0110135.

[49] Carroll, S. M., Field, G. B., and Jackiw, R., 1990, Phys. Rev. D, 41,

1231.

[50] Amelino-Camelia, G., Ellis, J. R., Mavramatos, N. E., Nanopoulos,

D. V., and Sarkar, S., 1998, Nature, 393, 763.

[51] Alfaro, J., Morales-Tecotl, H. A., and Urrutia, L. F., 2000, Phys. Rev.

Lett., 84, 2318.

[52] Choubey, S., and King, S. F., hep-ph/0207260.

[53] Amelino-Camelia, G., 2001, Phys. Lett. B, 510, 255; 2002, Int. J. Mod.

Phys., D, 11, 35; 2002, Int. J. Mod. Phys., D, 11, 1643.

[54] Magueijo, J., and Smolin, L., 2003, Phys. Rev. D, 67, 044017.

Graham Shore was awarded his PhD by the

University of Cambridge in 1978, following under-

graduate studies at the University of Edinburgh. He

subsequently held research appointments at Harvard

and Cornell, the Universities of Bern and Geneva,

and Imperial College, London where he was a

PPARC Advanced Fellow. He is a frequent visitor

to the European Particle Physics Laboratory at

CERN, Geneva where he has spent seven years as a

research associate. Since 1998, he has been Professor

of Theoretical Physics at the University of Wales,

Swansea.

521Quantum gravitational optics

Dow

nloa

ded

by [

Uni

vers

ity o

f W

este

rn O

ntar

io]

at 0

7:25

12

May

201

3