Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Optical analogues of black-hole horizons

Yuval Rosenberg
Department of Physics of Complex Systems,
Weizmann Institute of Science, Rehovot 761001, Israel.

April 14, 2020

Abstract
arXiv:2002.04216v2 [physics.optics] 12 Apr 2020

Hawking radiation is unlikely to be measured from a real black-hole, but can be tested in
laboratory analogues. It was predicted as a consequence of quantum mechanics and general
relativity, but turned out to be more universal. A refractive index perturbation produce an
optical analogue of the black-hole horizon and Hawking radiation that is made of light. We
discuss the central and recent experiments of the optical analogue, using hands-on physics.
We stress the roles of classical fields, negative frequencies, ’regular optics’ and dispersion.
Opportunities and challenges ahead are shortly discussed.

1 Introduction
Analogue systems are used to test theories of predicted phenomena that are hard to observe
directly [1]. Hawking [2, 3] predicted black-holes to thermally radiate from the quantum vac-
uum. Its measurement is unlikely, having a temperature inversely related to the black-hole
mass. Stellar and heavier black-holes’ radiation is much weaker than the cosmic microwave
background (CMB) fluctuations [4, 5]. Hypothetical microscopic primordial black-holes might
emit significant Hawking radiation [6], but it is highly model-dependent, with rates seriously
bounded by the lack of its observation [7–9].
Volovik and Unruh [10, 11] have shown that a transonic fluid flow is analogous to the space-
time geometry surrounding a black-hole, and should emit sound waves analogous to Hawking
radiation.
Wave kinematics control the Hawking process, which appears in the presence of an effective
horizon [12, 13]. Black-hole dynamics and Einstein equations of gravity are not an essential
requirement. The analogue of the event horizon is the surface separating the subsonic and
supersonic flows. The surface gravity determines the strength of Hawking radiation and is
analogous to the flow acceleration at the horizon. The horizon should also last long enough
such that the modes of Hawking radiation will have well defined frequencies [13].
Many analogues have been proposed [14, 15] with realisations using water waves [16–21],
Bose-Einstein condensates (BEC) [22–25], and optics [26–33]. The analogues gave new per-
spectives in both gravity and the analogue systems [14], extending Hawking’s theory to wave
dynamics in moving media. The universality of the Hawking effect shows it persists in real-
world scenarios, where complicated dynamics replace simplified and possibly fine-tuned theories.
Hawking’s original derivation seems questionable, since the radiation originates from infinitely
high frequencies, neglecting unknown high energy physics (known as the ’trans-Planckian prob-
lem’). The microscopic physics of the analogue systems is well known, allowing to test the role
of high frequencies in Hawking predictions [34, 35].
This paper discusses the optical analogues that use a refractive index perturbation to es-
tablish an artificial black-hole horizon. Extensive reviews of analogue gravity exist [14, 15], but
they pay little attention to optical analogues and substantial progress was made after their pub-
lication. Additional useful resources include Refs. [36–39]. This paper aims to present the main

1
experiments that transformed the optical analogues from simple ideas to established reality. It
aims to do so in the light of the recent developments, but in terms of hands-on physics. We
believe that experiments reach the state where their data may lead the way for new questions
and discoveries in the physics of Hawking radiation.
Section 2. briefly describes early proposals for optical analogues using slow light and why
they could not materialise. Transformation optics is also mentioned, where stationary dielectrics
are viewed to change the spatial parts of the metric. Section 3. explains the standard theory
of optical analogues discussed in this paper and the first demonstration of optical horizons.
Further measurements of the frequency shift at group velocity horizons are briefly discussed,
and additional related work is mentioned. Section 4. focuses on attempts in bulk optics rather
than optical fibres, stressing the roles of group-velocity and phase-velocity [40] horizons. Section
5. discusses the first measurements of negative frequencies in optics, a phenomenon closely
related to the Hawking effect. Its theory is also presented and shortly explained. Section 6.
considers additional interpretations of the optical horizon. Theory and experiments are shown
to directly relate the effect to cascaded four-wave mixing, but analysis in the time domain seems
to be unavoidable. The use of temporal analogues of reflection and refraction, and numerical
solutions are also mentioned. Section 7. discusses the first demonstration of stimulated Hawking
radiation in an optical analogue, and lessons learnt from it. Section 8. concludes the paper with
a brief outlook to the future.

2 Early attempts
Fresnel’s drag [41] was an ether-based theory of light propagation in transparent media, ’con-
firmed’ by Fizeau [42]. The drag effect is just a relativistic velocity addition [43], but Fresnel’s
wrong theory was based on correct intuition of velocity addition. The ether was replaced by the
space-time geometry and the quantum vacuum, which continued to have an intimate connec-
tion to moving media and produce puzzling phenomena [2,3,44–46]. Analogue gravity took this
connection a step further, but despite theoretical progress in the 1990’s, no practical realization
of analogue black-holes was suggested at that time [14].
Technology drove ideas in the right directions (optics and Bose-Einstein condensates) around
the year 2000 [47–51]. In optics, Leonhardt and Piwnicki [47–49] suggested to slow down light
such that its medium could be moved in super-luminal velocities and form a horizon. These ideas
used the technology of ’slow light’ [52, 53], where incredibly low group velocities are produced
using electromagnetically induced transparency (EIT, see Fig. 3a). However, the phase velocity
of ’slow light’ is fast, preventing the crucial formation of a phase velocity horizon (where the
media moves at the phase velocity of light. See below) [54]. Another problem with realising
these ideas is the narrow bandwidth of light that can be slowed [54], and severe absorption
around it [40]. The inevitable conclusion was to move the medium in relativistic velocities.
Despite missing key concepts of Hawking radiation, these ideas have pushed analogue grav-
ity forward and beyond the scope of relativity (or relativitists). They showed that an analogue
black-hole metric can be made in optics. Similarly, stationary dielectrics mimic spatial geome-
tries. This interpretation inspired transformation optics, and the development of meta-materials
technology extended the range of practical geometries [55–60].

3 Changing a reference frame


The groups of Leonhardt and König [26] started a new approach that follows a simple idea [61]:
light itself travels at the speed of light. A light pulse (also denoted pump) travels in a dielectric
medium with group velocity u . In the reference frame co-moving with the pulse (’the co-
moving frame’), the medium seems to flow in the opposite direction with velocity magnitude |u|
(Fig. 1b). Probe light of a different frequency differ in velocity due to dispersion. The nonlinear

2
(a)  !>0 (b)

!<0
u

Black-Hole
n = n0 + n2 I

Figure 1: Main optical analogue concepts. (a) A black-hole emit Hawking radiation from its
event horizon: positive norm waves escape outside, while negative norm waves drift towards
its singularity. The surface gravity κ represented by a falling weight determines the radiation’s
strength. (b) The medium seems to flow in the pulse co-moving frame with its group velocity u.
A pair of black-hole and white-hole horizons are formed due to nonlinear refraction index being
proportional to the pulse local intensity. Positive frequency probe light is shifted, and negative
frequency waves are formed at both horizons according to the dispersion relation (see Fig. 2).
The magnitude of the Hawking effect is determined by the pulse steepness, being analogous to
the surface gravity at the horizon (see text).

Kerr effect [62] slows down the probe upon interaction with the pulse1 – the refractive index
change as n = n0 + n2 I, where n0 is the linear refractive index, n2 is a parameter being typically
10−16 cm2 /W [62], and I is the light intensity. A group velocity horizon forms where the probe
group velocity obey vg = |u|, blocking the probe from further entering the pump. A black-hole
horizon is formed in the leading end of the pump, where the flow is directed into the pulse. Its
time reversal, a white-hole horizon with outward flow, is formed in the pump trailing end [26,63].
If the pump is slowly varying [13], the probe co-moving (and Doppler shifted) frequency,
 u
ω0 = γ 1 − n ω, (1)
c
−1/2
is conserved. Here γ = 1 − u2 /c2 is the Lorentz factor, c is the vacuum speed of light, n
is the refraction index, and ω is the probe frequency in the laboratory frame. Photon pairs of
Hawking radiation are produced at each horizon: one with positive ω 0 , and its negative partner
with −ω 0 [38, 64]. For positive ω, ω 0 is positive only if the probe phase velocity, vφ , is greater
than |u|. A phase velocity horizon forms where vφ = |u| so ω 0 = 0. The Hawking radiation
outgoing from the horizon mix incoming radiation of positive and negative frequencies [38].
Quantum mechanically, it is described by a Bogoliubov transformation

b̂± = αâ± + βâ†∓ ; |α|2 − |β|2 = 1, (2)

where the sign of the operators equals the sign of ω 0 . Time dependent annihilation and creation
operators for incoming modes, â± , â†∓ mix to form annihilation operators for outgoing modes,
b̂± . This extracts metric energy (pump energy) to amplify the radiation, thus spontaneously
creating outgoing radiation from incoming vacuum [33, 38]. The transformation parameters, α
and β, give the flux of Hawking radiation. When neglecting dispersion, the radiation effective
1
This has a similar effect to the flow acceleration in the fluid analogues, changing the probe velocity relative
to the flow.

3
Co-moving Frequency ω′ 0
SP Probe !probe

0
DW !pump
Pump PH
0 0
!pump NRR
0
!probe NHR

Lab Frequency ω
Figure 2: Schematic Doppler curve of the fibre optic analogue with phase matching conditions.
Frequency ω 0 at the reference frame co-moving with the pump pulse is plotted against the
laboratory frequency ω (Eq. 1). The pump pulse is in a local minima, where the dispersion is
anomalous. The phase matching condition of dispersive waves (DW) corresponds to conservation
0
of ωpump (upper dotted line). Negative frequency resonant radiation (NRR) conserve −ωpump 0

(lower dotted line. See also Fig. 3d,3f). Probe light (black-diamond) is red-shifted (white
0
diamond) at black-hole horizon, conserving its co-moving frequency ωprobe (upper solid line).
At the white-hole horizon the probe (now white diamond) is blue-shifted (black diamond), as
seen also in Fig. 3b. At the local maxima the group-velocity matches that of the pump, u.
Negative Hawking radiation (NHR) conserve −ωprobe0 (lower solid line. See also Fig. 3f). The
phase horizon (PH), where the phase velocity equals u, corresponds to ω 0 = 0.

temperature is proportional to the analogue of the surface gravity – the steepness of the pulse
(giving the probe velocity gradient in the co-moving frame [13, 38], which is related to the
refractive index (and pulse intensity) gradient or rate of change [26, 38, 63]).
In [26], an optical fibre called a photonic crystal fibre (PCF) provided both the desired
dispersion (Fig. 2) and nonlinearity. Its unique structure guided light in a very small region
– the fibre’s core [65, 66]. This increased the light intensity and the fibre’s nonlinear response.
Its nonlinear parameter was γ (ω0 ) = ω0 n2 /cAeff = 0.1W−1 m−1 at ω0 corresponding to 780 nm
wavelength, where Aeff is the fibre effective mode area and 0 is the vacuum permittivity [66].
The fibre’s structure was engineered to change its dispersion relation, n(w), to have two points
with matching group velocities [66]: one in a normal dispersion region, where n(w) is increasing;
and another in an anomalous dispersion region, where n(w) is decreasing. The anomalous
dispersion included the pump spectra, generated by a Ti:Sapphire mode-locked laser. Self-
phase modulation (SPM) due to the nonlinear refraction index counteracted the anomalous
dispersion and formed stable solitons [66].
The solitons in [26] created horizons. Continuous wave (CW) probe was added to the fibre,
being red-detuned from the frequency of matching group velocity (white diamond in Fig. 2).
The fast probe mainly interacted with the pump trailing end, gradually slowed down, and blue-
shifted (frequency up-conversion, see Fig. 3b). This pump-probe interaction is known as cross

4
phase modulation (XPM), and induces chirp that depends on the pump pulse shape [66]. If
δn = n2 I is large enough to form a horizon, the shifting probe become slower than the pulse
and they separate. In the co-moving frame the probe is reflected (Fig. 1b) while conserving its
0
co-moving frequency ωprobe , but not ω (Fig. 2). This reflection demonstrated the existence of
optical horizons [26].
Philbin et.al. [26] used 70 fs duration FWHM (full width at half maximum), 800 nm carrier
wavelength pulses with peak power of about 50 W. The PCF was 1.5 m long with core diameter
smaller than 2 µm. The CW probe was tunable around 1500 nm wavelength with 100 − 600 µW
power. Pump power was kept low to maintain a stable single soliton and reduce high-order
nonlinearities, such as the Raman effect [66]. Self-steepening was hoped to realize high Hawking
temperatures.
This set-up achieved a mere 10−3 percent efficiency for probe shifting. Only 10−4 of the CW
probe power was expected to interact with the pump over the 1.5 m fibre, and another order
of magnitude reduction was attributed to tunnelling of the probe through the narrow pump
barrier [26]. While clearly showing probe blue-shifting at the white-hole horizon, this set-up
could not produce detectable negative frequency partners. It was unclear whether they should
form around the phase velocity horizon, where ω 0 = 0, or around the supported frequency −ωin 0

(Fig. 2).
Choudhary and könig [67] reported probe red-shifting and studied its tunnelling. They
used similar experimental parameters to [26], with tuneable 50 fs pump pulses and visible CW
probe at 532 nm (to avoid the difficulty of synchronisation). Tunnelling was minimal for probe
detuning up to twice the soliton bandwidth (relative to the matching frequency), but limited
pump-probe interaction kept the conversion efficiency small. Tartara [68] derived both pump
and probe from the same 105 fs source and propagated them in a 1.1 m fibre, demonstrating
conversion efficiencies of tens of percents. An optical parametric amplifier produced the tuneable
probe.
The Raman effect cause the pump to decelerate, and was shown to only slightly change
the shifted probe spectra [69]. Shifting of dispersive waves [70, 71] and trapping the probe
light [72–76] were also related to the Raman effect. An optical ’black-hole laser’ that use both
the white and black-hole horizons was predicted and analysed [77–81]. The horizon dynamics
was also related to the formation of optical rogue waves and champion solitons [82–85], and the
ability to form all-optical transistors [86]. Similar ’front-induced transitions’ were studied in
other areas of photonics [87], and the quest for measuring analogue Hawking radiation continued.

4 Horizonless emissions
Faccio’s group realized optical phase velocity horizons, without a group velocity horizon, in
bulk optics [27, 88]. They constructed pulses of super-luminal group velocities, by using their
three dimensional nature [89]. A pulsed Bessel beam is one such example. It is composed of
infinitely many plane waves on a cone with a constant angle θ with the propagation axis. It can
be made with an axicon lens [89]. At the apex of the cone, the plane waves interfere to form
a bright spot that moves with super-luminal velocity that depends on θ. Belgiorno et al. [27]
used super-luminal pulses that travelled as narrow and powerful filaments inside fused silica.
Looking perpendicularly to their propagation direction, they detected radiation around the
phase velocity horizon – where the radiation phase velocity matched the pulses group velocity
(Fig. 3c). They could not measure along the propagation direction, because vast radiation was
produced by the powerful pulses there (peak intensity was as high as 1013 W/cm2 centred at
1055 nm wavelength).
Spurious radiation was a key issue even for observation at 90 degrees. Special care was
taken to tune the radiation into spectral windows of minimal noise. Correctly identifying all
the signals in such a set-up is another major challenge. Concerns were raised [90–92] and in-

5
depth analysis related the radiation to a horizonless super-luminal perturbation, possibly linked
to Hawking radiation [93, 94]. This work stressed the importance of a blocking group velocity
horizon to the Hawking effect. Thermal Hawking radiation was predicted for super-luminal
pulses in materials with linear dispersion at low energies (like diamond) [93, 94], but is yet to
be measured.
Leonhardt and Rosenberg [95] related the emission in [27] to the physics of Cherenkov
radiation2 in a surprising way. The pulse was modelled as a super-luminal light bullet [97], and
found to behave like a moving magnetic dipole that is predicted to emit Cherenkov radiation
with a discontinuous spectrum [98,99]. This discontinuity is exactly at the phase horizon – where
the effective dipole velocity equals the phase velocity of light and ω 0 = 0 (Fig. 2). Interferences
turn the discontinuity into a peak if the dipole is an extended object, like the light bullet.

5 The role of negative frequencies

Negative frequencies are an integral part of Hawking radiation and can also appear as dispersive
waves. Solitons emit dispersive waves (also known as resonant radiation) at a shifted frequency
when disturbed by higher order dispersion. Non-dispersive three dimensional light bullets emit
similar radiation [89]. Both emissions can be viewed to originate from self-scattering of the
pulse by its nonlinear refractive index barrier. Conservation of momentum dictates the emitted
frequency through a phase matching condition [66, 100]. In the reference frame co-moving
with the stable pulse, this condition becomes a conservation of energy, given by the pulse
0
co-moving frequency, ωpulse (Fig. 2). The frequency shifting at optical horizons extends this
picture to the scattering of probe light. The prediction of negative Hawking modes suggested
that negative dispersive waves should similarly appear. Rubino et al. [28] had measured this
negative-frequency resonant radiation (NRR) after it was neglected and disregarded for a long
time. This required very rapid (non-adiabatic) temporal changes of the pulse envelope.
Intense pulses created NRR through steep shocks in two ways [28]: 7 fs high order solitons
[66] of about 300 pJ energy and 800 nm carrier wavelength were placed in 5 mm PCFs; and
pulsed Bessel beams of 60 fs duration, about 20 µJ energy and 800 nm carrier wavelength were
sent through a 2 cm long bulk calcium fluoride (CaF2 ), as seen in Fig. 3d. Despite NRR was seen
both in optical fibres and in bulk, further work was needed to exclude other possible mechanisms
for the effect – contributions due to high order spatial modes, and possible interactions between
the complex pulse and additional radiation.
Rubino et al. [101] later used a relativistic scattering potential to directly explain NRR
formation, and Petev et al. [93] further related it to Hawking radiation. Conforti et al. [102]
extended the theory to pulses in materials of normal dispersion (where the pulse experience
dispersive broadening) and dominating second order nonlinearity, which effectively creates a
nonlinear refractive index. The same conditions for phase matching and non-adiabatic tempo-
ral evolution were found. Analytic derivation of the NRR [103] directly related it to interactions
between fields and their conjugates (or the mixing of positive and negative frequency modes).
It further strengthened the connection between NRR and Hawking radiation, which originate
from a Bogoliubov transformation that mixes creation and annihilation operators [38,64]. McLe-
naghan and König [104] studied NRR for different PCF lengths and input chirps, and inferred
UV propagation loss of ∼ 2dB/mm. We stress that the negative frequency (and negative norm)
of the Hawking modes beyond the horizon is key to the Hawking amplification process [38],
extracting energy from the background metric in accordance to a Bogoliubov transformation.

2
Not to be confused with dispersive waves [96]

6
6 ’It’s just optics’
Hawking radiation is a universal geometric effect that emerge due to conversion of modes at a
horizon, regardless of the microscopic physics that create the background space-time geometry
[13]. The same (generalized) derivation applies for both astrophysical black-holes and analogue
systems [13, 26, 105–110]. This established universality proved insightful for both gravity and
optics (arguably solving the trans-Planckian problem and discovering negative frequencies in
optics are two examples [14,15]). However, some researchers from both the optics and gravity (or
quantum field theory on curved space-times) communities are still not comfortable with analogue
Hawking radiation, claiming that the effect is ’just optics’ 3 . Such claims do not undermine
(analogue) Hawking radiation, but complete our understanding of it. It’s also possible to directly
explain the effect using the underlying microscopic physics – unknown quantum gravity for real
black-holes, and nonlinear optics for the optical analogues.
The classical dynamics at optical horizons can be captured using numerical solutions that
take into account the entire electric field (including negative frequencies), the dispersion relation
and all nonlinearities [66, 111–113]. These are extensions and alternatives for solving the usual
generalised nonlinear Schrödinger equation (GNLSE), which is used extensively in standard
descriptions of ultrafast nonlinear fibre optics [66, 100, 114].
Another approach directly relates the phenomena to discrete photonic interactions. It uses
cascaded four-wave mixing of discrete spectra, and takes the continuum limit for compari-
son [115, 116]. A CW probe and a pair of beating quasi-CWs (of 1 ns duration) positioned
symmetrically around the pump central frequency were being mixed continuously (Fig. 3e). At
each step of the cascaded process, the probe shifted by the pair’s detuning, and all three gener-
ated an equidistant frequency comb. A resonant amplification was produced at the frequency
0
corresponding to conservation of ωprobe . Similar amplification appeared at the dispersive wave
0
resonance, at ωpump . The experiments used low cascade orders (n = 5 − 7 for the probe shift
and about n = 15 for the dispersive waves) and were supported by numerical analysis. This
picture shows how energy is being transferred from the pump to the shifting probe, by effectively
absorbing pump photons of one frequency and emitting at another. The cascaded four wave
mixing also details the back reaction mechanism, where the pump undergoes spectral recoil.
The studies [115, 116] addressed the relevant phase matching conditions, but the efficiency
of the cascaded process was found only in [117], for dispersive waves. Remarkably, it showed
that the known concepts from horizon physics are crucial ingredients even when the pump
is made of beating quasi-CWs that form a frequency comb: the mixing was analysed using
(temporal) soliton fission dynamics, being efficient only when the frequency spacing between
the CWs effectively generated compressing (non-adiabatic) high order solitons. The frequency-
domain analysis was intractable for high cascade orders. The maximally compressed beat cycle
corresponded to about 102 fs duration FWHM in the efficient regime, of high cascading orders.
Conversion efficiencies of 10−4 over a 100 m fibre were reported.
The generation of NRR in media with quadratic nonlinearity was also related to a cascading
process [102], which must also be the case for cubic materials. However, negative frequencies
were not demonstrated yet using a beating CW pump. It would be interesting to test how the
cascading mixing behaves as it reaches negative frequencies, past the phase velocity horizon.
Temporal analogues of reflection and refraction [118, 119] are also used to explain the fre-
quency shifts at optical horizons. It compares the frequency change due to XPM to the wave
number change during refraction. The reflection at the horizon is analogous to total internal
reflection. The tunnelling through the refractive index barrier is compared to frustrated total
internal reflection, where an evanescent wave extends beyond the pulse and tunnels through.
3
The author have personally encountered such allegations from both communities, and believe that they
partially originate from not separating ’gravity’, predicted to create the black-hole event horizon, and ’kinematics’,
which are responsible for the Hawking effect once a horizon is formed [13].

7
The optical horizon can be seen as a temporal beam splitter for probe light (acting also as an
amplifier when considering negative frequencies as well).
Comparing the mode mixing and squeezing of parametric amplification [120] to the Hawking
effect is also useful [38]. The origin of the two phenomena is completely different, and even in the
optical analogue, simple down conversion obey different phase matching conditions that conserve
total lab frequency [62, 66] and cannot explain Hawking radiation. However, the two share
intuition and the Bogoliubov transformation between incoming and outgoing modes. In both
cases a pump creates ’signal’ and ’idler’ waves: the vacuum spontaneously generates quantum
emissions, while stimulating the effect with classical light amplifies the effect for specific modes.

7 Observing stimulated Hawking radiation

Spontaneous Hawking radiation originate from the quantum vacuum. Classical laser fields can
replace the vacuum and stimulate the effect. Drori et al. [33] observed stimulated negative
Hawking radiation in an optical analogue (Fig. 3f). The idea of [26] was used, but with more
suitable pump pulses and a pulsed probe. Analysis of the fibre dispersion (schematically seen in
Fig. 2) and simulations of the experiment allowed to optimise the experimental parameters. The
pump was a high order few-cycle soliton that compressed and collapsed due to soliton fission [66].
Its peak power reached a few hundred kilo-watts, with 800 nm central wavelength. Very non-
adiabatic dynamics generated NRR in the mid UV (Fig. 3f), at the negative of the pump
co-moving frequency, −ωpump0 . This ensured the formation of steep refractive index variations,
increasing the analogous surface gravity and reducing the probability of probe tunnelling. The
probe was derived from the pump’s laser, and was generated through cascaded Raman scattering
in a meter long PCF. Negative prechirp was used to enhance the Raman induced frequency
shift [121], whose wavelength was continuously tuned up to 1650 nm by varying the input power.
Peak powers were around 1 kW.
Efficient and broad-band probe frequency shifts at the horizon accompanied negative Hawk-
ing radiation in the mid-UV, conserving the probe co-moving frequency ωprobe 0 and its conjugate
0
−ωprobe , respectively. Since the fibre had strong dispersion, with different notions for phase and
group velocity horizons, the spectrum was not Planckian [33, 64, 107]. Varying ωprobe 0 , both
signals shifted according to theory, supporting the correct interpretation of the negative Hawk-
ing radiation. By this, also the interpretation of the NRR was supported, and its analytic
theory [103] was generalised to include the Hawking process [113]. Linear relation was verified
between the probe power and that of the negative Hawking signals up to a point where the
signal saturated. This saturation was related to a back-reaction of the probe on the pulse,
which is a prerequisite for Hawking radiation (since its energy is drawn from the pump, or the
black-hole that curves the metric). It also slightly reduced the magnitude of the NRR (Fig. 3f).
The rate of spontaneous Hawking emission was estimated based on the magnitude of the
stimulated effect [33]. It was found to be too low for measurement in the system used, calling to
design an improved setup. The predicted spontaneous effect is minuscule and is overwhelmed by
the noise in the system, which for the negative Hawking radiation is dominated by fluctuations
of the overlapping NRR. The multi-mode nature of the PCF in the UV [66] was found to reduce
the observed signal power by up to four orders of magnitude.
The experiment [33] showed how robust the Hawking effect is: it appeared despite the
extreme nonlinear dynamics of the collapsing soliton. Since time in the co-moving frame is
related to the propagation distance along the fibre, rapid variations in the pump profile do not
violate the required slow evolution of the metric as long as they appear over a length scale
much larger than the Hawking radiation wavelength [13]. As such, pump deceleration due to
the Raman effect could be accounted by simply correcting its central frequency.

8
9
peakthatshiftstoshorterwavelengthswithincreasinginput
Bessel pulses and thus underlining the absence of any dicted to depend on " n which in
f uorescence or(a)other possible emission signals when(b)the " n ¼ n2I of the pulse intensity.
energy. )
a
(
T his process has
RIP v does not satisfy Eq. (1). Theother |four
been
3
described
1curves show
in detail
f t gives n2 ¼ 2:8)[20] 0:5*in10'

Normalized power density [ppm]


similar
the emitted1.01 conditions
spectra andinput
for four different is energies
a directof the manifestation
good
w/o pulse
of the
[dBm] agreement with12the tabul
with pulse [dBm]
' 16 2
2 10 cm =W [12,25]. Therefore

Power density [dBm]


Bessel f lament, as indicated in the f gure. A clear photon
formation of a steep |
shock front on
theory fitthe trailing edge
difference [ppm]
Refractive index

9
emission is registered in thewavelength window |1 predicted also an agreement at the quanti

Absorption
byof
Eq. the 1 pump
(1). Moreover pulse.
weverif ed that As energy
theemitted is increased,
radiation measurementsthe and shock
the
6
model
was unpolarized (data not shown) thus further supporting radiation emission.
front
the steepens
interpretation and theemission.
of a spontaneous spectral peak shifts
In addition, In Fig. toward shorter
4 we present 3additional

wavelengths.
the measurements
0.99
Between
clearly show that the 15 and
bandwidth #20
of the "emissionJ input froma spontaneous f la
energy, 0
a
emission is increasing -10 with the
0 input10energy. The Bessel
0
1480 1490 1500 1510
Detuning [MHz]
loosely
Wavelength focusing
[nm] a 50 $ J Gauss
(c) silica sample. The spontaneous
60 (d)
lead to theformation of af lamen
Photoelectron counts

Spectral density
to the Bessel f lament in the sen

[arb. units]
40
by a very high-intensity peak th
distances, thus creating a strong
20
may be expected to excite Hawk
300 fashion500 to that
Wavelength [nm]
700 illustrated abov
0

750 850 950


Wavelength [nm]
(e) (f)
10 Pump
Counts [kHz]

Pump & Probe


Signal
5

0
226 228 230 232 234
Wavelength [nm]

Figure 3: Central measurements. (a) Schematic absorption (dotted line) and refractive index
(solid line) for generating ’slow light’ using electromagnetically induced transparency (EIT), as
the inset depicts. The refractive index has small but rapid variation at the narrow resonance.
FIG. 3 (color online). (a) Spectra generated by a Bessel f la-
The slow group velocity is inversely related to this steep change [52, 53]. (b) Frequency blue-
ment. Five different curves are shown corresponding to a refer-
shift of probe light at fibre optical white-hole event horizon. 70 fs pump solitons were used in
ence spectrum obtained with a Gaussian pulse (black line) and
1.5 m long nonlinear PCF. A small fraction of CW probe at ω1 is converted to ω2 , conserving
ndicated Bessel energies (in $ J). 0 Dashed lines are guides for
its co-moving frequency, ωprobe . From [26]; reprinted with permission from 2008 c AAAS. (c)
the eye. The solid line connects the spectral peaks, highlighting
Horizonless emissions from a Bessel filament creating a super-luminal refractive index pertur-
the +40 nm shift with increasing energy, in close agreement
bation
with the in bulk45
predicted silica [27, 93].
nm shift. (b)The radiationatwas
Bandwidths FWHMseen around
and the phase velocity horizon (where
0
RIP "ωn = 0) and
versus increased
input energywith pump energy
and intensity. µJ).linear
Solid(inline: The fwavelength
t shift (black line) and broad-
ening was in agreement
" n ¼ n2I with n2 ¼ 2:8* 10 with
' 16 the
2 increased
cm =W. nonlinear refractive index. From [27]; reprinted with
c
permission from 2010 APS (d) Negative-frequency resonant radiation (NRR) in bulk CaF2 .
16 µJ Bessel pump beam at 800 nm created dispersive-waves (around 600 nm) and NRR (around
350 nm) under highly non-adiabatic temporal evolution. From [28]; reprinted with permission
c
from 2012 APS. (e) Comparison of frequency translation induced by a pump soliton (orange
curve) and a pair of quasi-CW fields (purple curve). Probe light is being blue-shifted (’Idler’) in
a similar way in both cases, relating the effect to cascaded four-wave mixing. Similar dispersive-
waves (DW) are also seen. From [115]; reprinted with permission from 2014 c Springer Nature.
(f ) Negative frequency stimulated Hawking radiation. A compressing high-order few-cycle soli-
ton [33] generated horizons and NRR (green dotted-dashed curve). When a pulsed probe at
1450 nm interacted with the black-hole horizon, it red shifted (not shown) and produced neg-
ative Hawking radiation that overlapped the NRR (black solid curve). Subtracting the NRR
revealed the UV Hawking signal (blue curve with 2σ error bars). From [33]; reprinted with
permission from 2019 c APS.
8 Beyond the horizon?
Analogue gravity in optics has come a long way: from erroneous visionary ideas, to careful
analysis of experimental demonstrations. It drove new research directions and understandings
in optics. The reality of optical horizons and negative frequencies became clear, and multiple
optical interpretations made their physics more tangible. The crucial role of dispersion was
stressed, determining the Hawking spectrum (Fig. 2). A group velocity horizon is blocking
the radiation modes and allow their efficient conversion. A phase velocity horizon marks the
support of non-trivial negative frequency modes, that allow for the Hawking amplification. The
optical and other analogues, especially in water waves and BECs, provide extensive insights for
gravity through concrete real-world examples. The universality of the Hawking effect separates
it from gravity and high-energy physics, and emphasises the central role of classical fields to the
process. This hints to more possibilities in gravity [15] and in optics [122], even before going
fully quantum.
The main challenge for observing the spontaneous (quantum) effect in optics is its low power.
Noise from spurious radiation makes matters worse. Using knowledge from the stimulated
effect [33] could help overcome these challenges.
The demonstrated robustness of the Hawking effect is calling for new experiments to lead the
way. Our increased confidence in the interpretation of the results allows to further separate from
the traditional scheme of Hawking radiation, developing bolder questions and ideas. Optical
technologies allow to realise wild ideas and to study them with high precision. We call for more
cross-fertilisation between research of the different black-hole analogues, which might open new
horizons in these fields.

9 Funding
Fellowship of the Sustainability and Energy Research Initiative (SAERI) program, the Weiz-
mann Institute of Science; European Research Council; and the Israel Science Foundation.

10 Acknowledgements
I am grateful for discussions and comments from David Bermudez, Jonathan Drori and Ulf
Leonhardt. I acknowledge valuable discussions with the participants of the scientific meeting
’The next generation of analogue gravity experiments’ at the Royal Society (London, December
2019), and thank its organizers: Maxime Jacquet, Silke Weinfurtner and Friedrich König.

References
[1] I. M. Georgescu, S. Ashhab, and F. Nori, “Quantum simulation,” Rev. Mod. Phys, vol. 86,
no. 1, p. 153, 2014.

[2] S. W. Hawking, “Black hole explosions?,” Nature, vol. 248, no. 5443, pp. 30–31, 1974.

[3] S. W. Hawking, “Particle creation by black holes,” Commun. Math. Phys, vol. 43, no. 3,
pp. 199–220, 1975.

[4] D. Samtleben, S. Staggs, and B. Winstein, “The cosmic microwave background for pedes-
trians: A review for particle and nuclear physicists,” Annu. Rev. Nucl. Part. Sci., vol. 57,
pp. 245–283, 2007.

[5] The temperature of Hawking radiation from a black-hole of solar mass is about 60 nK,
where CMB temperature fluctuations are about 100 µK.
10
[6] F. Halzen, E. Zas, J. MacGibbon, and T. Weekes, “Gamma rays and energetic particles
from primordial black holes,” Nature, vol. 353, no. 6347, p. 807, 1991.

[7] D. Alexandreas, G. Allen, D. Berley, S. Biller, R. Burman, M. Cavalli-Sforza, C. Chang,


M. Chen, P. Chumney, D. Coyne, et al., “New limit on the rate-density of evaporating
black holes,” Phys. Rev. Lett, vol. 71, no. 16, p. 2524, 1993.

[8] C. Fichtel, D. Bertsch, B. Dingus, J. Esposito, R. Hartman, S. Hunter, G. Kanbach,


D. Kniffen, Y. Lin, J. Mattox, et al., “Search of the energetic gamma-ray experiment tele-
scope (egret) data for high-energy gamma-ray microsecond bursts,” Astrophys. J, vol. 434,
pp. 557–559, 1994.

[9] E. Linton, R. Atkins, H. Badran, G. Blaylock, P. Boyle, J. Buckley, K. Byrum, D. Carter-


Lewis, O. Celik, Y. Chow, et al., “A new search for primordial black hole evaporations
using the whipple gamma-ray telescope,” J. Cosmol. Astropart. Phys., vol. 2006, no. 01,
p. 013, 2006.

[10] G. E. Volovik, The universe in a helium droplet, vol. 117. Oxford University Press on
Demand, 2003.

[11] W. G. Unruh, “Experimental black-hole evaporation?,” Phys. Rev. Lett, vol. 46, no. 21,
p. 1351, 1981.

[12] M. Visser, “Acoustic black holes: horizons, ergospheres and hawking radiation,” Class.
Quantum Gravity, vol. 15, no. 6, p. 1767, 1998.

[13] M. Visser, “Essential and inessential features of hawking radiation,” Int. J. Mod. Phys.
D, vol. 12, no. 04, pp. 649–661, 2003.

[14] C. Barcelo, S. Liberati, and M. Visser, “Analogue gravity,” Living Rev. Relativ, vol. 14,
no. 1, p. 3, 2011.

[15] C. Barceló, “Analogue black-hole horizons,” Nat. Phys, vol. 15, no. 3, pp. 210–213, 2019.

[16] G. Rousseaux, C. Mathis, P. Maı̈ssa, T. G. Philbin, and U. Leonhardt, “Observation of


negative-frequency waves in a water tank: a classical analogue to the hawking effect?,”
New J. Phys, vol. 10, no. 5, p. 053015, 2008.

[17] G. Jannes, R. Piquet, P. Maı̈ssa, C. Mathis, and G. Rousseaux, “Experimental demonstra-


tion of the supersonic-subsonic bifurcation in the circular jump: A hydrodynamic white
hole,” Phys. Rev. E, vol. 83, no. 5, p. 056312, 2011.

[18] S. Weinfurtner, E. W. Tedford, M. C. Penrice, W. G. Unruh, and G. A. Lawrence, “Mea-


surement of stimulated hawking emission in an analogue system,” Phys. Rev. Lett, vol. 106,
no. 2, p. 021302, 2011.

[19] L.-P. Euvé, F. Michel, R. Parentani, and G. Rousseaux, “Wave blocking and partial
transmission in subcritical flows over an obstacle,” Phys. Rev. D, vol. 91, no. 2, p. 024020,
2015.

[20] L.-P. Euvé, F. Michel, R. Parentani, T. G. Philbin, and G. Rousseaux, “Observation of


noise correlated by the hawking effect in a water tank,” Phys. Rev. Lett, vol. 117, no. 12,
p. 121301, 2016.

[21] T. Torres, S. Patrick, A. Coutant, M. Richartz, E. W. Tedford, and S. Weinfurtner,


“Rotational superradiant scattering in a vortex flow,” Nat. Phys, vol. 13, no. 9, p. 833,
2017.

11
[22] O. Lahav, A. Itah, A. Blumkin, C. Gordon, S. Rinott, A. Zayats, and J. Steinhauer,
“Realization of a sonic black hole analog in a bose-einstein condensate,” Phys. Rev. Lett,
vol. 105, no. 24, p. 240401, 2010.

[23] J. Steinhauer, “Observation of self-amplifying hawking radiation in an analogue black-hole


laser,” Nat. Phys, vol. 10, no. 11, pp. 864–869, 2014.

[24] J. Steinhauer, “Observation of quantum hawking radiation and its entanglement in an


analogue black hole,” Nat. Phys, vol. 12, no. 10, p. 959, 2016.

[25] J. R. M. de Nova, K. Golubkov, V. I. Kolobov, and J. Steinhauer, “Observation of ther-


mal hawking radiation and its temperature in an analogue black hole,” Nature, vol. 569,
no. 7758, p. 688, 2019.

[26] T. G. Philbin, C. Kuklewicz, S. Robertson, S. Hill, F. König, and U. Leonhardt, “Fiber-


optical analog of the event horizon,” Science, vol. 319, no. 5868, pp. 1367–1370, 2008.

[27] F. Belgiorno, S. L. Cacciatori, M. Clerici, V. Gorini, G. Ortenzi, L. Rizzi, E. Rubino,


V. G. Sala, and D. Faccio, “Hawking radiation from ultrashort laser pulse filaments,”
Phys. Rev. Lett, vol. 105, no. 20, p. 203901, 2010.

[28] E. Rubino, J. McLenaghan, S. Kehr, F. Belgiorno, D. Townsend, S. Rohr, C. Kuklewicz,


U. Leonhardt, F. König, and D. Faccio, “Negative-frequency resonant radiation,” Phys.
Rev. Lett, vol. 108, no. 25, p. 253901, 2012.

[29] M. Elazar, V. Fleurov, and S. Bar-Ad, “All-optical event horizon in an optical analog of
a laval nozzle,” Phys. Rev. A, vol. 86, no. 6, p. 063821, 2012.

[30] H. S. Nguyen, D. Gerace, I. Carusotto, D. Sanvitto, E. Galopin, A. Lemaı̂tre, I. Sagnes,


J. Bloch, and A. Amo, “Acoustic black hole in a stationary hydrodynamic flow of micro-
cavity polaritons,” Phys. Rev. Lett, vol. 114, no. 3, p. 036402, 2015.

[31] R. Bekenstein, R. Schley, M. Mutzafi, C. Rotschild, and M. Segev, “Optical simulations of


gravitational effects in the newton–schrödinger system,” Nat. Phys, vol. 11, no. 10, p. 872,
2015.

[32] R. Bekenstein, Y. Kabessa, Y. Sharabi, O. Tal, N. Engheta, G. Eisenstein, A. J. Agranat,


and M. Segev, “Control of light by curved space in nanophotonic structures,” Nat. Pho-
tonics, vol. 11, no. 10, p. 664, 2017.

[33] J. Drori, Y. Rosenberg, D. Bermudez, Y. Silberberg, and U. Leonhardt, “Observation of


stimulated hawking radiation in an optical analogue,” Phys. Rev. Lett, vol. 122, no. 1,
p. 010404, 2019.

[34] T. Jacobson, “Black-hole evaporation and ultrashort distances,” Phys. Rev. D, vol. 44,
no. 6, p. 1731, 1991.

[35] W. G. Unruh, “Sonic analogue of black holes and the effects of high frequencies on black
hole evaporation,” Phys. Rev. D, vol. 51, no. 6, p. 2827, 1995.

[36] D. B. Francesco, L. C. Sergio, and F. Daniele, Hawking radiation: from astrophysical black
holes to analogous systems in lab. World Scientific, 2018.

[37] D. Faccio, F. Belgiorno, S. Cacciatori, V. Gorini, S. Liberati, and U. Moschella, Analogue


gravity phenomenology: analogue spacetimes and horizons, from theory to experiment,
vol. 870. Springer, 2013.

12
[38] U. Leonhardt, Essential quantum optics: from quantum measurements to black holes.
Cambridge University Press, 2010.

[39] W. Unruh and R. Schützhold, Quantum analogues: from phase transitions to black holes
and cosmology, vol. 718. Springer, 2007.

[40] P. W. Milonni, Fast light, slow light and left-handed light. CRC Press, 2004.

[41] A. Fresnel, “Lettre d’augustin fresnel à françois arago sur l’influence du mouvement ter-
restre dans quelques phénomènes d’optique,” Ann. Chim. Phys., vol. 9, pp. 57–66, 1818.

[42] A. Fizeau, “Sur les hypothèses relatives à l’éther lumineux, et sur une expérience qui
parait démontrer que le mouvement des corps change la vitesse avec laquelle la lumière
se propage dans leur intérieur,” C. R. Acad. Sci, vol. 33, pp. 349–355, 1851.

[43] W. Rindler, Relativity: Special, General, and Cosmological. Oxford University Press,
2006.

[44] S. A. Fulling, “Nonuniqueness of canonical field quantization in riemannian space-time,”


Phys. Rev. D, vol. 7, no. 10, p. 2850, 1973.

[45] P. C. Davies, “Scalar production in schwarzschild and rindler metrics,” J. Phys. A, vol. 8,
no. 4, p. 609, 1975.

[46] W. G. Unruh, “Notes on black-hole evaporation,” Phys. Rev. D, vol. 14, no. 4, p. 870,
1976.

[47] U. Leonhardt and P. Piwnicki, “Optics of nonuniformly moving media,” Phys. Rev. A,
vol. 60, no. 6, p. 4301, 1999.

[48] U. Leonhardt and P. Piwnicki, “Relativistic effects of light in moving media with extremely
low group velocity,” Phys. Rev. Lett, vol. 84, no. 5, p. 822, 2000.

[49] U. Leonhardt, “A laboratory analogue of the event horizon using slow light in an atomic
medium,” Nature, vol. 415, no. 6870, p. 406, 2002.

[50] L. J. Garay, J. Anglin, J. I. Cirac, and P. Zoller, “Sonic analog of gravitational black holes
in bose-einstein condensates,” Phys. Rev. Lett, vol. 85, no. 22, p. 4643, 2000.

[51] L. J. Garay, J. Anglin, J. I. Cirac, and P. Zoller, “Sonic black holes in dilute bose-einstein
condensates,” Phys. Rev. A, vol. 63, no. 2, p. 023611, 2001.

[52] L. V. Hau, S. E. Harris, Z. Dutton, and C. H. Behroozi, “Light speed reduction to 17


metres per second in an ultracold atomic gas,” Nature, vol. 397, no. 6720, p. 594, 1999.

[53] M. M. Kash, V. A. Sautenkov, A. S. Zibrov, L. Hollberg, G. R. Welch, M. D. Lukin,


Y. Rostovtsev, E. S. Fry, and M. O. Scully, “Ultraslow group velocity and enhanced
nonlinear optical effects in a coherently driven hot atomic gas,” Phys. Rev. Lett, vol. 82,
no. 26, p. 5229, 1999.

[54] W. Unruh and R. Schützhold, “On slow light as a black hole analogue,” Phys. Rev. D,
vol. 68, no. 2, p. 024008, 2003.

[55] U. Leonhardt, “Optical conformal mapping,” Science, vol. 312, no. 5781, pp. 1777–1780,
2006.

[56] J. B. Pendry, D. Schurig, and D. R. Smith, “Controlling electromagnetic fields,” Science,


vol. 312, no. 5781, pp. 1780–1782, 2006.

13
[57] U. Leonhardt and T. G. Philbin, “General relativity in electrical engineering,” New J.
Phys, vol. 8, no. 10, p. 247, 2006.

[58] H. Chen, C. T. Chan, and P. Sheng, “Transformation optics and metamaterials,” Nature
materials, vol. 9, no. 5, p. 387, 2010.

[59] U. Leonhardt and T. Philbin, Geometry and light: the science of invisibility. Courier
Corporation, 2010.

[60] L. Xu and H. Chen, “Conformal transformation optics,” Nat. Photonics, vol. 9, no. 1,
p. 15, 2015.

[61] U. Leonhardt and F. König, “Invention disclosure,” 2005. Research and Enterprise ser-
vices, University of St Andrews.

[62] R. W. Boyd, Nonlinear optics. Elsevier, 2003.

[63] F. Belgiorno, S. Cacciatori, G. Ortenzi, L. Rizzi, V. Gorini, and D. Faccio, “Dielectric


black holes induced by a refractive index perturbation and the hawking effect,” Phys. Rev.
D, vol. 83, no. 2, p. 024015, 2011.

[64] U. Leonhardt and S. Robertson, “Analytical theory of hawking radiation in dispersive


media,” New J. Phys, vol. 14, no. 5, p. 053003, 2012.

[65] P. Russell, “Photonic crystal fibers,” Science, vol. 299, no. 5605, pp. 358–362, 2003.

[66] G. P. Agrawal, Nonlinear Fiber Optics. Academic Press, 2019.

[67] A. Choudhary and F. König, “Efficient frequency shifting of dispersive waves at solitons,”
Opt. Express, vol. 20, no. 5, pp. 5538–5546, 2012.

[68] L. Tartara, “Frequency shifting of femtosecond pulses by reflection at solitons,” IEEE J.


Quantum Electron, vol. 48, no. 11, pp. 1439–1442, 2012.

[69] S. Robertson and U. Leonhardt, “Frequency shifting at fiber-optical event horizons: the
effect of raman deceleration,” Phys. Rev. A, vol. 81, no. 6, p. 063835, 2010.

[70] S. Wang, A. Mussot, M. Conforti, A. Bendahmane, X. Zeng, and A. Kudlinski, “Optical


event horizons from the collision of a soliton and its own dispersive wave,” Phys. Rev. A,
vol. 92, no. 2, p. 023837, 2015.

[71] A. Bendahmane, A. Mussot, M. Conforti, and A. Kudlinski, “Observation of the stepwise


blue shift of a dispersive wave preceding its trapping by a soliton,” Opt. Express, vol. 23,
no. 13, pp. 16595–16601, 2015.

[72] N. Nishizawa and T. Goto, “Pulse trapping by ultrashort soliton pulses in optical fibers
across zero-dispersion wavelength,” Opt. Lett, vol. 27, no. 3, pp. 152–154, 2002.

[73] N. Nishizawa and T. Goto, “Characteristics of pulse trapping by use of ultrashort soliton
pulses in optical fibers across the zero-dispersion wavelength,” Opt. Express, vol. 10,
no. 21, pp. 1151–1159, 2002.

[74] A. V. Gorbach and D. V. Skryabin, “Light trapping in gravity-like potentials and expan-
sion of supercontinuum spectra in photonic-crystal fibres,” Nat. Photonics, vol. 1, no. 11,
p. 653, 2007.

[75] S. Hill, C. Kuklewicz, U. Leonhardt, and F. König, “Evolution of light trapped by a


soliton in a microstructured fiber,” Opt. Express, vol. 17, no. 16, pp. 13588–13601, 2009.

14
[76] W. Wang, H. Yang, P. Tang, C. Zhao, and J. Gao, “Soliton trapping of dispersive waves in
photonic crystal fiber with two zero dispersive wavelengths,” Opt. Express, vol. 21, no. 9,
pp. 11215–11226, 2013.

[77] S. Corley and T. Jacobson, “Black hole lasers,” Phys. Rev. D, vol. 59, no. 12, p. 124011,
1999.

[78] U. Leonhardt and T. G. Philbin, “Black hole lasers revisited,” in Quantum Analogues:
From Phase Transitions to Black Holes and Cosmology, pp. 229–245, Springer, 2007.

[79] D. Faccio, T. Arane, M. Lamperti, and U. Leonhardt, “Optical black hole lasers,” Class.
Quantum Gravity, vol. 29, no. 22, p. 224009, 2012.

[80] J. L. Gaona-Reyes and D. Bermudez, “The theory of optical black hole lasers,” Ann.
Phys., vol. 380, pp. 41–58, 2017.

[81] D. Bermudez and U. Leonhardt, “Resonant hawking radiation as an instability,” Class.


Quantum Gravity, vol. 36, no. 2, p. 024001, 2018.

[82] D. R. Solli, C. Ropers, P. Koonath, and B. Jalali, “Optical rogue waves,” Nature, vol. 450,
no. 7172, pp. 1054–1057, 2007.

[83] A. Demircan, S. Amiranashvili, C. Brée, C. Mahnke, F. Mitschke, and G. Steinmeyer,


“Rogue events in the group velocity horizon,” Sci. Rep, vol. 2, p. 850, 2012.

[84] A. Demircan, S. Amiranashvili, C. Brée, C. Mahnke, F. Mitschke, and G. Steinmeyer,


“Rogue wave formation by accelerated solitons at an optical event horizon,” Appl. Phys.
B, vol. 115, no. 3, pp. 343–354, 2014.

[85] S. Pickartz, U. Bandelow, and S. Amiranashvili, “Adiabatic theory of solitons fed by


dispersive waves,” Phys. Rev. A, vol. 94, no. 3, p. 033811, 2016.

[86] A. Demircan, S. Amiranashvili, and G. Steinmeyer, “Controlling light by light with an


optical event horizon,” Phys. Rev. Lett, vol. 106, no. 16, p. 163901, 2011.

[87] M. A. Gaafar, T. Baba, M. Eich, and A. Y. Petrov, “Front-induced transitions,” Nat.


Photonics, vol. 13, no. 11, pp. 737–748, 2019.

[88] E. Rubino, F. Belgiorno, S. Cacciatori, M. Clerici, V. Gorini, G. Ortenzi, L. Rizzi, V. Sala,


M. Kolesik, and D. Faccio, “Experimental evidence of analogue hawking radiation from
ultrashort laser pulse filaments,” New J. Phys, vol. 13, no. 8, p. 085005, 2011.

[89] D. Faccio, A. Couairon, and P. Di Trapani, “Conical waves, filaments, and nonlinear
filamentation optics,” Aracne, 2007.

[90] R. Schützhold and W. G. Unruh, “Comment on: Hawking radiation from ultrashort laser
pulse filaments,” arXiv preprint arXiv:1012.2686, 2010.

[91] W. Unruh and R. Schützhold, “Hawking radiation from ?phase horizons? in laser fila-
ments?,” Phys. Rev. D, vol. 86, no. 6, p. 064006, 2012.

[92] S. Liberati, A. Prain, and M. Visser, “Quantum vacuum radiation in optical glass,” Phys.
Rev. D, vol. 85, no. 8, p. 084014, 2012.

[93] M. Petev, N. Westerberg, D. Moss, E. Rubino, C. Rimoldi, S. Cacciatori, F. Belgiorno, and


D. Faccio, “Blackbody emission from light interacting with an effective moving dispersive
medium,” Phys. Rev. Lett, vol. 111, no. 4, p. 043902, 2013.

15
[94] S. Finazzi and I. Carusotto, “Spontaneous quantum emission from analog white holes in
a nonlinear optical medium,” Phys. Rev. A, vol. 89, no. 5, p. 053807, 2014.
[95] U. Leonhardt and Y. Rosenberg, “Cherenkov radiation of light bullets,” Phys. Rev. A,
vol. 100, no. 6, p. 063802, 2019.
[96] N. Akhmediev and M. Karlsson, “Cherenkov radiation emitted by solitons in optical
fibers,” Phys. Rev. A, vol. 51, no. 3, p. 2602, 1995.
[97] Y. Silberberg, “Collapse of optical pulses,” Opt. Lett, vol. 15, no. 22, pp. 1282–1284, 1990.
[98] I. Frank, “Doppler effect in a refractive medium,” Bull. Russ. Acad. Sci.: Phys., vol. 6,
no. 3, p. 2, 1942.
[99] I. Frank, “Vavilov-cherenkov radiation for electric and magnetic multipoles,” Phys.-
Uspekhi, vol. 27, no. 10, pp. 772–785, 1984.
[100] J. M. Dudley, G. Genty, and S. Coen, “Supercontinuum generation in photonic crystal
fiber,” Rev. Mod. Phys, vol. 78, no. 4, p. 1135, 2006.
[101] E. Rubino, A. Lotti, F. Belgiorno, S. Cacciatori, A. Couairon, U. Leonhardt, and D. Fac-
cio, “Soliton-induced relativistic-scattering and amplification,” Sci. Rep, vol. 2, p. 932,
2012.
[102] M. Conforti, N. Westerberg, F. Baronio, S. Trillo, and D. Faccio, “Negative-frequency
dispersive wave generation in quadratic media,” Phys. Rev. A, vol. 88, no. 1, p. 013829,
2013.
[103] M. Conforti, A. Marini, T. X. Tran, D. Faccio, and F. Biancalana, “Interaction between
optical fields and their conjugates in nonlinear media,” Opt. Express, vol. 21, no. 25,
pp. 31239–31252, 2013.
[104] J. McLenaghan and F. König, “Few-cycle fiber pulse compression and evolution of negative
resonant radiation,” New J. Phys, vol. 16, no. 6, p. 063017, 2014.
[105] S. Finazzi and I. Carusotto, “Quantum vacuum emission in a nonlinear optical medium
illuminated by a strong laser pulse,” Phys. Rev. A, vol. 87, no. 2, p. 023803, 2013.
[106] M. Jacquet and F. König, “Quantum vacuum emission from a refractive-index front,”
Phys. Rev. A, vol. 92, no. 2, p. 023851, 2015.
[107] D. Bermudez and U. Leonhardt, “Hawking spectrum for a fiber-optical analog of the event
horizon,” Phys. Rev. A, vol. 93, no. 5, p. 053820, 2016.
[108] M. F. Linder, R. Schützhold, and W. G. Unruh, “Derivation of hawking radiation in
dispersive dielectric media,” Phys. Rev. D, vol. 93, no. 10, p. 104010, 2016.
[109] M. Jacquet, Negative Frequency at the horizon: theoretical study and experimental reali-
sation of analogue gravity physics in dispersive optical media. Springer, 2018.
[110] M. J. Jacquet and F. Koenig, “Analytical description of quantum emission in optical
analogues to gravity,” arXiv preprint arXiv:1908.02060, 2019.
[111] S. Amiranashvili, “Hamiltonian framework for short optical pulses,” in New Approaches
to Nonlinear Waves, pp. 153–196, Springer, 2016.
[112] D. Bermudez, “Propagation of ultra-short higher-order solitons in a photonic crystal
fiber,” in Journal of Physics: Conference Series, vol. 698, p. 012017, IOP Publishing,
2016.

16
[113] R. Aguero, Y. Rosenberg, U. Leonhardt, and D. Bermudez forthcoming 2020.

[114] D. V. Skryabin and A. V. Gorbach, “Colloquium: Looking at a soliton through the prism
of optical supercontinuum,” Rev. Mod. Phys, vol. 82, no. 2, p. 1287, 2010.

[115] K. E. Webb, M. Erkintalo, Y. Xu, N. G. Broderick, J. M. Dudley, G. Genty, and S. G.


Murdoch, “Nonlinear optics of fibre event horizons,” Nat. Commun, vol. 5, p. 4969, 2014.

[116] M. Erkintalo, Y. Xu, S. Murdoch, J. Dudley, and G. Genty, “Cascaded phase matching
and nonlinear symmetry breaking in fiber frequency combs,” Phys. Rev. Lett, vol. 109,
no. 22, p. 223904, 2012.

[117] K. Webb, M. Erkintalo, Y. Xu, G. Genty, and S. Murdoch, “Efficiency of dispersive wave
generation from a dual-frequency beat signal,” Opt. Lett, vol. 39, no. 20, pp. 5850–5853,
2014.

[118] B. Plansinis, W. Donaldson, and G. Agrawal, “What is the temporal analog of reflection
and refraction of optical beams?,” Phys. Rev. Lett, vol. 115, no. 18, p. 183901, 2015.

[119] B. W. Plansinis, W. R. Donaldson, and G. P. Agrawal, “Cross-phase-modulation-induced


temporal reflection and waveguiding of optical pulses,” JOSA B, vol. 35, no. 2, pp. 436–
445, 2018.

[120] C. Gerry, P. Knight, and P. L. Knight, Introductory quantum optics. Cambridge university
press, 2005.

[121] Y. Rosenberg, J. Drori, D. Bermudez, and U. Leonhardt, “Boosting few-cycle soliton


self-frequency shift using negative prechirp,” Opt. Express, vol. 28, pp. 3107–3115, Feb
2020.

[122] U. Leonhardt, “On cosmology in the laboratory,” Philos. Trans. R. Soc. A, vol. 373,
no. 2049, p. 20140354, 2015.

17

You might also like