We have seen that one of the main aims of research in analogue models of gravity is the possibility of
simulating semiclassical gravity phenomena, such as the Hawking radiation effect or cosmological particle
production. In this sense systems characterised by a high degree of quantum coherence, very cold
temperatures, and low speeds of sound offer the best test field. One could reasonably hope to manipulate
these systems to have Hawking temperatures on the order of the environment temperature
( 100 nK) [48
]. Hence it is not surprising that in recent years Bose–Einstein condensates
(BECs) have become the subject of extensive study as possible analogue models of general
relativity [231
, 232
, 45
, 48
, 47
, 195
, 194
].
Let us start by very briefly reviewing the derivation of the acoustic metric for a BEC system, and show
that the equations for the phonons of the condensate closely mimic the dynamics of a scalar field in a
curved spacetime. In the dilute gas approximation, one can describe a Bose gas by a quantum field
satisfying
We would also like to highlight that in relative terms, the approximation by which one neglects the
quartic terms in the dispersion relation gets worse as one moves closer to a horizon where
. The non-dimensional parameter that provides this information is defined by
Indeed, with hindsight, the fact that the group velocity goes to infinity for large was pre-ordained:
After all, we started from the generalised nonlinear Schrödinger equation, and we know what its
characteristic curves are. Like the diffusion equation the characteristic curves of the Schrödinger
equation (linear or nonlinear) move at infinite speed. If we then approximate this generalised
nonlinear Schrödinger equation in any manner, for instance by linearization, we cannot change the
characteristic curves: For any well-behaved approximation technique, at high frequency and
momentum we should recover the characteristic curves of the system we started with. However, what
we certainly do see in this analysis is a suitably large region of momentum space for which the concept
of the effective metric both makes sense, and leads to finite propagation speed for medium-frequency
oscillations.
In [191] an analogue model based on a relativistic BEC was studied. We summarise here the main
results. The Lagrangian density for an interacting relativistic scalar Bose field
may be written as
The field can be written as a classical field (the condensate) plus perturbation:
A detailed discussion of the different regimes would be inappropriately long for this review; it can be
found in [191]. The results are summarised in Table 1. Note that
plays the role of the chemical
potential for the relativistic BEC. One of the most remarkable features of this model is that it is a
condensed matter system that interpolates between two different Lorentz symmetries, one at low energy and
a different Lorentz symmetry at high energy.
Gapless
|
Gapped | |||
|
|
|||
![]() |
![]() |
|||
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
|||
![]() |
![]() |
|||
Finally, it is also possible to recover an acoustic metric for the massless (phononic) perturbations of the
condensate in the low momentum limit ():
Helium is one of the most fascinating elements provided by nature. Its structural richness confers on helium
a paradigmatic character regarding the emergence of many and varied macroscopic properties from the
microscopic world (see [660] and references therein). Here, we are interested in the emergence
of effective geometries in helium, and their potential use in testing aspects of semiclassical
gravity.
Helium four, a bosonic system, becomes superfluid at low temperatures (2.17 K at vapour pressure). This superfluid behaviour is associated with condensation in the vacuum state of a macroscopically large number of atoms. A superfluid is automatically an irrotational and inviscid fluid, so, in particular, one can apply to it the ideas worked out in Section 2. The propagation of classical acoustic waves (scalar waves) over a background fluid flow can be described in terms of an effective Lorentzian geometry: the acoustic geometry. However, in this system one can naturally go considerably further, into the quantum domain. For long wavelengths, the quasiparticles in this system are quantum phonons. One can separate the classical behaviour of a background flow (the effective geometry) from the behaviour of the quantum phonons over this background. In this way one can reproduce, in laboratory settings, different aspects of quantum field theory over curved backgrounds. The speed of sound in the superfluid phase is typically on the order of cm/sec. Therefore, at least in principle, it should not be too difficult to establish configurations with supersonic flows and their associated ergoregions.
Helium three, the fermionic isotope of helium, in contrast, becomes superfluid at much lower
temperatures (below 2.5 milli-K). The reason behind this rather different behaviour is the pairing of
fermions to form effective bosons (Cooper pairing), which are then able to condense. In the 3He-A phase,
the structure of the fermionic vacuum is such that it possesses two Fermi points, instead of the more typical
Fermi surface. In an equilibrium configuration one can choose the two Fermi points to be located at
(in this way, the z-axis signals the direction of the angular momentum of the
pairs). Close to either Fermi point the spectrum of quasiparticles becomes equivalent to that of Weyl
fermions. From the point of view of the laboratory, the system is not isotropic, it is axisymmetric.
There is a speed for the propagation of quasiparticles along the z-axis,
, and a
different speed,
, for propagation perpendicular to the symmetry axis. However,
from an internal observer’s point of view this anisotropy is not “real”, but can be made to
disappear by an appropriate rescaling of the coordinates. Therefore, in the equilibrium case, we
are reproducing the behaviour of Weyl fermions over Minkowski spacetime. Additionally, the
vacuum can suffer collective excitations. These collective excitations will be experienced by
the Weyl quasiparticles as the introduction of an effective electromagnetic field and a curved
Lorentzian geometry. The control of the form of this geometry provides the sought for gravitational
analogy.
Apart from the standard way to provide a curved geometry based on producing nontrivial flows, there is
also the possibility of creating topologically nontrivial configurations with a built-in nontrivial geometry.
For example, it is possible to create a domain-wall configuration [327, 326
] (the wall contains the z-axis)
such that the transverse velocity
acquires a profile in the perpendicular direction (say
along the x-axis) with
passing through zero at the wall (see Figure 11
). This particular
arrangement could be used to reproduce a black-hole–white-hole configuration only if the soliton is
set up to move with a certain velocity along the x-axis. This configuration has the advantage
that it is dynamically stable, for topological reasons, even when some supersonic regions are
created.
A third way in which superfluid helium can be used to create analogues of gravitational configurations is
the study of surface waves (or ripplons) on the interface between two different phases of 3He
[657, 659
]. In particular, if we have a thin layer of 3He-A in contact with another thin layer of
3He-B, the oscillations of the contact surface “see” an effective metric of the form [657
, 659
]
The advantage of using surface waves instead of bulk waves in superfluids is that one could create
horizons without reaching supersonic speeds in the bulk fluid. This could alleviate the appearance of
dynamical instabilities in the system, that in this case are controlled by the strength of the interaction of
the ripplons with bulk degrees of freedom [657, 659].
The geometrical interpretation of the motion of light in dielectric media leads naturally to conjecture that the use of flowing dielectrics might be useful for simulating general relativity metrics with ergoregions and black holes. Unfortunately, these types of geometry require flow speeds comparable to the group velocity of the light. Since typical refractive indexes in non-dispersive media are quite close to unity, it is then clear that it is practically impossible to use them to simulate such general relativistic phenomena. However recent technological advances have radically changed this state of affairs. In particular the achievement of controlled slowdown of light, down to velocities of a few meters per second (or even down to complete rest) [617, 338, 96, 353, 506, 603, 565], has opened a whole new set of possibilities regarding the simulation of curved-space metrics via flowing dielectrics.
But how can light be slowed down to these “snail-like” velocities? The key effect used to achieve this takes the name of Electromagnetically Induced Transparency (EIT). A laser beam is coupled to the excited levels of some atom and used to strongly modify its optical properties. In particular one generally chooses an atom with two long-lived metastable (or stable) states, plus a higher energy state that has some decay channels into these two lower states. The coupling of the excited states induced by the laser light can affect the transition from a lower energy state to the higher one, and hence the capability of the atom to absorb light with the required transition energy. The system can then be driven into a state where the transitions between each of the lower energy states and the higher energy state exactly cancel out, due to quantum interference, at some specific resonant frequency. In this way the higher-energy level has null averaged occupation number. This state is hence called a “dark state”. EIT is characterised by a transparency window, centered around the resonance frequency, where the medium is both almost transparent and extremely dispersive (strong dependence on frequency of the refractive index). This in turn implies that the group velocity of any light probe would be characterised by very low real group velocities (with almost vanishing imaginary part) in proximity to the resonant frequency.
Let us review the most common setup envisaged for this kind of analogue model. A more detailed
analysis can be found in [383]. One can start by considering a medium in which an EIT window is opened
via some control laser beam which is oriented perpendicular to the direction of the flow. One then
illuminates this medium, now along the flow direction, with some probe light (which is hence perpendicular
to the control beam). This probe beam is usually chosen to be weak with respect to the control beam, so
that it does not modify the optical properties of the medium. In the case in which the optical properties of
the medium do not vary significantly over several wavelengths of the probe light, one can neglect the
polarization and can hence describe the propagation of the latter with a simple scalar dispersion
relation [390
, 211]
It is easy to see that in this case the group and phase velocities differ
So even for small refractive indexes one can get very low group velocities, due to the large dispersion in the transparency window, and in spite of the fact that the phase velocity remains very near toAt resonance we have
We can now generalise the above discussion to the case in which our highly dispersive medium flows with a characteristic velocity profile Several comments are in order concerning the metric (297). First of all, it is clear that, although more
complicated than an acoustic metric, it will still be possible to cast it into the Arnowitt–Deser–Misner-like
form [627]
In any case, the existence of this ADM form already tells us that an ergoregion will always appear once
the norm of the effective velocity exceeds the effective speed of light (which for slow light is approximately
, where
can be extremely large due to the huge dispersion in the transparency window
around the resonance frequency
). However, a trapped surface (and hence an optical black
hole) will form only if the inward normal component of the effective flow velocity exceeds the
group velocity of light. In the slow light setup so far considered such a velocity turns out to be
.
The realization that ergoregions and event horizons can be simulated via slow light may lead one to the
(erroneous) conclusion that this is an optimal system for simulating particle creation by gravitational fields.
However, as pointed out by Unruh in [470, 612
], such a conclusion would turn out to be over-enthusiastic.
In order to obtain particle creation through “mode mixing”, (mixing between the positive and negative
norm modes of the incoming and outgoing states), an inescapable requirement is that there must
be regions where the frequency of the quanta as seen by a local comoving observer becomes
negative.
In a flowing medium this can, in principle, occur thanks to the tilting of the dispersion relation
due to the Doppler effect caused by the velocity of the flow Equation (293); but this also tells
us that the negative norm mode must satisfy the condition
, but this can be
satisfied only if the velocity of the medium exceeds
, which is the phase velocity of the
probe light, not its group velocity. This observation suggests that the existence of a “phase
velocity horizon” is an essential ingredient (but not the only essential ingredient) in obtaining
Hawking radiation. A similar argument indicates the necessity for a specific form of “group
velocity horizon”, one that lies on the negative norm branch. Since the phase velocity in the
slow light setup we are considering is very close to
, the physical speed of light in vacuum,
not very much hope is left for realizing analogue particle creation in this particular laboratory
setting.
However, it was also noticed by Unruh and Schützhold [612] that a different setup for slow light might
deal with this and other issues (see [612
] for a detailed summary). In the setup suggested by these authors
there are two strong-background counter-propagating control beams illuminating the atoms. The field
describing the beat fluctuations of this electromagnetic background can be shown to satisfy, once the
dielectric medium is in motion, the same wave equation as that on a curved background. In this particular
situation the phase velocity and the group velocity are approximately the same, and both can be made
small, so that the previously discussed obstruction to mode mixing is removed. So in this new setup it is
concretely possible to simulate classical particle creation such as, e.g., super-radiance in the presence of
ergoregions.
Nonetheless, the same authors showed that this does not open the possibility for a simulation of quantum particle production (e.g., Hawking radiation). This is because that effect also requires the commutation relations of the field to generate the appropriate zero-point energy fluctuations (the vacuum structure) according to the Heisenberg uncertainty principle. This is not the case for the effective field describing the beat fluctuations of the system we have just described, which is equivalent to saying that it does not have a proper vacuum state (i.e., analogue to any physical field). Hence, one has to conclude that any simulation of quantum particle production is precluded.
In addition to the studies of slow light in fluids, there has now been a lot of work done on slow light in a
fibre-optics setting [505, 504
, 64
, 63
], culminating in recent experimental detection of photons
apparently associated with a phase-velocity horizon [66
]. The key issue here is that the Kerr
effect of nonlinear optics can be used to change the refractive index of an optical fibre, so that
a “carrier” pulse of light traveling down the fibre carries with it a region of high refractive
index, which acts as a barrier to “probe” photons (typically at a different frequency). If the
relative velocities of the “carrier” pulse and “probe” are suitably arranged then the arrangement
can be made to mimic a black-hole–white-hole pair. This system is described more fully in
Section 6.4.
The quantum analogue models described above all have an underlying discrete structure: namely the atoms
they are made of. In abstract terms one can also build an analogue model by considering a quantum field on
specific lattice structures representing different spacetimes. In [310, 149
, 322] a falling-lattice black-hole
analogue was put forward, with a view to analyzing the origin of Hawking particles in black-hole
evaporation. The positions of the lattice points in this model change with time as they follow
freely falling trajectories. This causes the lattice spacing at the horizon to grow approximately
linearly with time. By definition, if there were no horizons, then for long wavelengths compared
with the lattice spacing one would recover a relativistic quantum field theory over a classical
background. However, the presence of horizons makes it impossible to analyze the field theory only
in the continuum limit, it becomes necessary to recall the fundamental lattice nature of the
model.
A very interesting addition to the catalogue of analogue systems is the graphene (see, for example, these
reviews [127, 339]). Although graphene and some of its peculiar electronic properties have been known
since the 1940s [672], only recently has it been specifically proposed as a system with which to probe
gravitational physics [153
, 152
]. Graphene (or mono-layer graphite) is a two-dimensional lattice of carbon
atoms forming a hexagonal structure (see Figure 13
). From the perspective of this review, its most
important property is that its Fermi surface has two independent Fermi points (see Section 4.2.2 on
helium). The low-energy excitations around these points can be described as massless Dirac
fields in which the light speed is substituted by a “sound” speed
about 300 times smaller:
From this perspective graphene can be used to investigate ultra-relativistic phenomena such as the Klein paradox [339]. On the other hand, graphene sheets can also acquire curvature. A nonzero curvature can be produced by adding strain fields to the sheet, imposing a curved substrate, or by introducing topological defects (e.g., some pentagons within the hexagonal structure) [669]. It has been suggested that, regarding the electronic properties of graphene, the sheet curvature promotes the Dirac equation to its curved space counterpart, at least on the average [153, 152]. If this proves to be experimentally correct, it will make graphene a good analogue model for a diverse set of spacetimes. This set, however, does not include black-hole spacetimes, as the curvatures mentioned above are purely spatial and do not affect the temporal components of the metric.
http://www.livingreviews.org/lrr-2011-3 |
Living Rev. Relativity 14, (2011), 3
![]() This work is licensed under a Creative Commons License. E-mail us: |