You are on page 1of 8

ARTICLE

https://doi.org/10.1038/s41467-019-12815-0 OPEN

Topological control of extreme waves


Giulia Marcucci 1,2*, Davide Pierangeli1,2, Aharon J. Agranat3, Ray-Kuang Lee 4, Eugenio DelRe1,2 &
Claudio Conti 1,2

From optics to hydrodynamics, shock and rogue waves are widespread. Although they appear
as distinct phenomena, transitions between extreme waves are allowed. However, these have
1234567890():,;

never been experimentally observed because control strategies are still missing. We intro-
duce the new concept of topological control based on the one-to-one correspondence
between the number of wave packet oscillating phases and the genus of toroidal surfaces
associated with the nonlinear Schrödinger equation solutions through Riemann theta func-
tions. We demonstrate the concept experimentally by reporting observations of supervised
transitions between waves with different genera. Considering the box problem in a focusing
photorefractive medium, we tailor the time-dependent nonlinearity and dispersion to explore
each region in the state diagram of the nonlinear wave propagation. Our result is the first
realization of topological control of nonlinear waves. This new technique casts light on shock
and rogue waves generation and can be extended to other nonlinear phenomena.

1 Department of Physics, University Sapienza, Piazzale Aldo Moro 5, 00185 Rome, Italy. 2 Institute for Complex Systems, Via dei Taurini 19, 00185 Rome,

Italy. 3 Applied Physics Department, Hebrew University of Jerusalem, 91904 Jerusalem, Israel. 4 Institute of Photonics Technologies, National Tsing Hua
University, Hsinchu 300, Taiwan. *email: giulia.marcucci@uniroma1.it

NATURE COMMUNICATIONS | (2019)10:5090 | https://doi.org/10.1038/s41467-019-12815-0 | www.nature.com/naturecommunications 1


ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12815-0

I
n 1967 Gardner, Greene, Kruskal, and Miura developed a
a mathematical method—the inverse scattering transform 240 Shock phase Breathers Soliton gas 8
(IST)1—disclosing the inner features of nonlinear waves in 220 7
hydrodynamics, plasma physics, nonlinear optics and many other 200
6
physical systems2–4. According to IST, one also predicts the 180
5
periodical regeneration of the initial state, as in the Fermi-Pasta-

x (µm)
160 I/I0
Ulam-Tsingou recurrence5,6. 4
The nonlinear Schrödinger equation (NLSE)7 is a cornerstone 140
3
of IST for detailing dispersive phenomena, such as dispersive 120
2
shock waves (DSWs)8–10, rogue waves (RWs)11–14, and shape 100
1
invariant solitons15–17. DSWs regularize catastrophic dis- 80
continuities by means of rapid oscillations18–22. RWs are giant 0 5 10 15 20 25 30 35 40 45 50
0

disturbances appearing and disappearing abruptly in a nearly t (second)


constant background23–34. Solitons are particle-like dispersion-
free wave packets that can form complex interacting assemblies, b 0.5
g=0
ranging from crystals to gases15,16,33,35–37. 0.45 Dispersive
DSWs, RWs, and soliton gases (SGs) are related phenomena, 0.4
shock waves

and all appear in paradigmatic nonlinear evolutions, such as the 0.35 g=1
box problem for the focusing NLSE38–43. However, for the box Rogue
0.3
problem in the small-dispersion NLSE, IST becomes unfeasible. waves

(W0, t)
0.25
In this extreme regime, the problem can be tackled by the so-
called finite-gap theory40,44. It turns out that extreme waves are 0.2 g=2
g >>2
described in terms of one single mathematical entity, the Rie- 0.15
mann theta function, and classified by a topological index, the 0.1
genus g (see Fig. 1). In nonlinear wave theory, g represents the 0.05 Soliton gas
number of oscillating phases and evolves during light propaga- 0
tion: “single phase” DSWs have g ¼ 1, RWs have g  2 and SGs 20 60 100 140
W0 (µm)
have g >> 2. This creates a fascinating connection between
extreme waves and topology. Indeed, the same genus g allows a Fig. 1 Topological classification of extreme waves. a Final states of the wave
topological classification of surfaces, to distinguish, for examples, for a fixed initial waist W 0 ¼ 100 μm showing the generation of focusing
a torus and sphere (Fig. 1). The question lies open if this elegant dispersive shock waves (g ¼ 1), rogue waves (g  2), and a soliton gas
mathematical classification of extreme waves can inspire new (g >> 2) after different time intervals in a photorefractive material (see
applications. Can it modify the basic paradigm by which the text). b Phase diagram reporting the final states in terms of the parameter ϵ
asymptotic evolution of a wave is encoded in its initial shape, and the initial beam waist. Transitions occur by fixing waist and varying ϵ
opening the way to controlling extreme waves, from lasers to or, equivalently, the observation time t. Different surfaces displayed in
earthquakes? proximity of the various wave profiles, corresponding to the different
Here, inspired by the topological classification, we propose and regions in the phase diagram, outline the link between the topological
demonstrate the use of topological indices to control the gen- classification of extreme waves in terms of the genus g and the topological
eration of extreme waves with varying genera g 41. We consider classification of toroidal Riemann surfaces (for a sphere, g ¼ 0, for a torus,
the NLSE box problem where, according to recent theoretical g ¼ 1, etc.)
results40, light experiences various dynamic phases during pro-
pagation, distinguished by different genera. In particular, for high which exhibits some of the most interesting dynamic phases in
values of a nonlinearly scaled propagation distance ζ, one has nonlinear wave propagation40,45. The initial evolution presents
g  ζ. By continuously varying ζ, we can change g and explore all the formation of two wave trains counterpropagating that reg-
the possible dynamic phases (see Fig. 1, where ζ is given in terms ularize the box discontinuities. These wave trains are single-phase
of the observation time t, detailed below). We experimentally test DSWs (g ¼ 1). Their two wavefronts superimpose in the central
this approach in photorefractive materials, giving evidence of an part of the box (see Fig. 1a)–occurring at ζ ¼ ζ 0 :¼ 2pl ffiffi2q – and
unprecedented control of nonlinear waves, which allows the first
observation of the transition from focusing DSWs to RWs. generate a breather lattice of genus g ¼ 2, a two-phase quasi-
periodic wave resembling an ensemble of Akhmediev breathers
Results (ABs)13,28. Since both the ξ and ζ periods increase with ζ, the
Time-dependent spatial box problem. We consider the NLSE oscillations at ξ ’ 0 become locally approximated by Peregrine
solitons (PSs)13,46–48. At long propagation distances ζ >> ζ 0 , the
ϵ2 2 wave train becomes multi-phase and generates a SG with g  ζ.
{ϵ∂ζ ψ þ ∂ ψ þ jψj2 ψ ¼ 0; ð1Þ
2 ξ In Fig. 1a, we report the wave dynamics in physical units, as we
make specific reference to our experimental realization of the
where ψ ¼ ψðξ; ζÞ is the normalized complex field envelope, ζ is NLSE box problem for spatial optical propagation in photo-
the propagation coordinate, ξ is the transverse coordinate and refractive media (PR). In these materials, the optical nonlinearity
ϵ > 0 is the dispersion parameter. We take a rectangular barrier is due to the time-dependent accumulation of free carriers
as initial condition that induces a time-varying low-frequency electric field. Through
 the electro-optic effect, the charge accumulation results into
q for jξj ≤ l
ψðξ; 0Þ ¼ ; ð2Þ a time-varying nonlinearity. The corresponding time-profile can
0 elsewhere be controlled by an external applied voltage and the intensity
that is, a box of finite height q > 0, length 2l > 0, and genus level49–51. These features enable to experimentally implement our
g ¼ 0. In our work, we fix q ¼ l ¼ 1. Equation (1) with (2) is topological control technique. In PR, Eq. (1) describes an optical
known as the NLSE box problem, or the dam break problem, beam with complex amplitude Aðz; x; tÞ and intensity I ¼ jAj2

2 NATURE COMMUNICATIONS | (2019)10:5090 | https://doi.org/10.1038/s41467-019-12815-0 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12815-0 ARTICLE

through the transformation (see Methods) correspondingly. The genus time-dependence is sketched in
z 2x A Fig. 1a. The output wave profile depends on its genus content,
ζ¼ ; ξ¼ ; ψ ¼ pffiffiffiffi ; ð3Þ which varies with t.
ϵzD W0 I0 Following the theoretical approach in ref. 40, the two separatrix
πn W 2 equations divide the evolution diagram in Fig. 1a into three
with W 0 the initial beam waist along x-direction, z D ¼ 2λ 0 0
the
2δn0 I different areas: the flat box plateau with genus g ¼ 0, the lateral
diffraction length, n ¼ n0 þ I f ðtÞ the refractive index, δn0 > 0 counterpropagating DSWs with genus g ¼ 1, and the RWs after
S
the nonlinear coefficient, I S the saturation intensity, I 0 the initial the DSW-collision point (corresponding to the separatrices
intensity. For PR intersection) with genus g ¼ 2. The two separatrices (dashed
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi lines in Fig. 1a) have equations
λ IS
ϵ¼ ; ð4Þ W0
πW 0 2n0 δn0 I 0 f ðtÞ x ¼ x0 ± ðt  t 0 Þ ¼ x0 ± vðt  t 0 Þ; ð5Þ
2t 0
namely, the dispersion is modulated by the time-dependent τI n W 2
crystal response function f ðtÞ ¼ 1  expðt=τ Þ, with the satura- with ðt 0 ; x0 Þ the DSW-collision point, t 0 ’ 64IS δn
0 0
2 , and x 0 given
0 0L
tion time τ fixed by the input power and the applied voltage5. by the central position of the box. It turns out that the shock
velocity is
Genus control. For a given propagation distance L (the length of
the photorefractive crystal), the genus of the final state is deter- W0 32δn0 L2
v¼ ¼ P; ð6Þ
mined by the detection time t, which determines ϵ, ζ ¼ ϵzL , and g, 2t 0 I S n0 W 20 U 0 τ
D

a I/I0
0 1 2 3 4 5 6 7 8
300 0.3

250 0.25

200 0.2
x (µm)

150 0.15

100 0.1

50 0.05

0 0
0 20 40 60 80 100 120
t (second)

b 5 c 10 d 10
= 0.081 = 0.055
I/I0

I/I0

I/I0

5 5

0 0 0
100 200 100 150 200 100 150 200
x (µm) x (µm) x (µm)

e 10 f 10 g 10
= 0.029 = 0.023 = 0.021
I/I0

I/I0

I/I0

5 5 5

0 0 0
100 150 200 100 150 200 100 150 200
x (µm) x (µm) x (µm)

Fig. 2 Controlling the extreme wave genus. a Numerical simulation of the control of the final state after a propagation distance L ¼ 2:5 mm for an initial
beam waist W 0 ¼ 140 μm (I0 ¼ U PW ¼ 0:38 ´ 105 W/m2). Axis x represents the beam transverse direction, axis t the time of output detection. b Initial
0 0
beam intensity: a super-Gaussian wave centered at x ¼ 150 μm of height I0 and width W 0 . c, d Focusing dispersive shock waves occurrence: c represents
the beam intensity at t ¼ 5 s, when the wave breaking has just occurred, so two lateral intense wave trains regularize the box discontinuity and start
to travel towards the beam central part; d the beam intensity at t ¼ 11 s, which exhibits the two counterpropagating DSWs reaching the center x ¼ 150 μm.
e–g Akhmediev breathers and Peregrine solitons generation: beam intensity at e t ¼ 49 s, f t ¼ 98 s, and g t ¼ 120 s, after the two dispersive shock waves
superposition and the formation of Akhmediev breathers with period increasing with t. Since a Peregrine soliton is an Akhmediev breather with an infinite
period, increasing t is tantamount to generating central intensity peaks, locally described by Peregrine solitons

NATURE COMMUNICATIONS | (2019)10:5090 | https://doi.org/10.1038/s41467-019-12815-0 | www.nature.com/naturecommunications 3


ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12815-0

a Pump laser
b
CW laser

Mask
3

/0 (×102)
OL
CCD
CL 2

E
KLTN
1
0 50 100 150 200
y P0 (µW)
x

c I/I0
Shock phase Breathers Soliton gas
2

g=1

200

g=0 g=2
x (µm)

g >>2
1
4
100 g=1

I/I0
2

0
76 106.5 137
x (µm)
0 0
0 20 40 60 80 100 120
t (second)

Fig. 3 Experimental demonstration of the extreme wave genus control. a Experimental setup. A CW laser is made a quasi-one-dimensional wave by a
cylindrical lens (CL), then a tunable mask shapes it as a box. Light propagates in a pumped photorefractive KLTN crystal, it is collected by a microscope
objective and the optical intensity is detected by a CCD camera. The inset shows an example of the detected input intensity distribution (scale bar is
50 μm). b Normalized shock velocity [v0 ¼ L=t, L ¼ 2:5 mm, t ¼ ð30 ± 2Þ s], measured through the width of the oscillation tail at fixed time, versus input
power. The blue squares are the experimental data, while the dashed pink line is the linear fit. c Experimental observation of optical intensity I=I0 for an
initial beam waist W 0 ¼ 140 μm. Axis x represents the beam profile, transverse to propagation, collected by the CCD camera, while axis t is time of CCD
camera detection. Output presents a first dispersive-shock-wave phase, a transition to a phase presenting Akhmediev breather structures and, at long
times, a generation of a soliton gas. The inset is an exemplary wave intensity profile detected at t ¼ 63 s (dotted blue line), along with the theoretical
Akhmediev breather profile

proportional to the input power, as experimentally demonstrated box, is shown in Supplementary Information (Suppl. Fig. 1).
and detailed below. When the DSWs superimpose, ABs are generated (Fig. 2e). From
Equation (5) expresses the genus time-dependence for its first the analytical NLSE solutions for the focusing dam break pro-
three values g ¼ 0; 1; 2. It allows designing the waveshape, before blem40, we see that ABs have ξ-period increasing with ζ. More-
the experiment, by associating a specific combination of the over, one finds that ∂t ζ > 0, therefore the period in the
topological indices, and to predict the detection time corre- x-direction must increase with time, and central peaks appear
sponding to the target topology. In other words, by properly upon evolution. These peaks are well approximated by PSs, for
choosing the experimental conditions, we can predict the large t, as confirmed by Fig. 2f, g.
occurrence of a given extreme wave by using the expected genus The occurrence of RWs in the large box regime is proved also
g. According to Eq. (4), we use time t and initial waist W 0 to by statistical analysis, illustrated in Supplementary Information
vary ϵ. The accessible states are outlined in the phase diagram in (Suppl. Fig. 2h, i).
Fig. 1b, in terms of ϵ and W 0 . Choosing W 0 ¼ 100 μm as in Figure 3 shows the experimental observation of the controlled
Fig. 1a, by varying t one switches from DSWs to RWs, and then dynamics simulated in Fig. 2. Figure 3a sketches the experimental
to SGs. setup, detailed in Methods. A quasi-one-dimensional box-shaped
beam propagates in a photorefractive crystal, and the optical
intensity distribution is detected at different times. The observa-
Supervised transition from shock to rogue waves. The case tions of shock velocities and beam propagation for W 0 ¼ 140 μm
W 0 ¼ 140 μm is illustrated in Fig. 2a by numerical simulations. are reported in Fig. 3b, c, respectively. In Fig. 3c, we see an initial
The two focusing DSWs and the SG are visible at the beginning DSW phase that evolves into a train of large amplitude waves. In
and at the end of temporal evolution, respectively (see phase this regime, we identify a breather-like structure (ABs, inset in
diagram in Fig. 1b). As soon as an initial super-Gaussian wave Fig. 3c) that evolves into a SG at large propagation time. The
(Fig. 2b, see Methods) starts to propagate, two DSWs appear on DSW phase is investigated varying the input power. We find a
the beam borders (Fig. 2c) and propagate towards the beam linear increasing behavior of the shock velocity when increasing
central part (Fig. 2d). Experimental proof of the genuine non- the power (Fig. 3b), as predicted by Eq. (6). The shock velocity is
linear nature of the beam evolution at this regime, not due to proportional to the distance between the two counterpropagating
modulation instability arising from noise in the central part of the DSWs at a fixed time. We measured the width Δx of the plateau

4 NATURE COMMUNICATIONS | (2019)10:5090 | https://doi.org/10.1038/s41467-019-12815-0 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12815-0 ARTICLE

a I/I0
0 1 2 3 4 5 6 7 8
220 4

3.5
200

3
180

2.5
160
x (µm)

140
1.5

120
1

100 0.5

80 0
0 20 40 60 80 100 120
t (second)

b 8 c d
= 0.74 = 0.36
6 5 5
I/I0

I/I0
I/I0

2 0 0
140 160 140 160
0
140 160 x (µm) x (µm)
x (µm)

e f g
= 0.35 = 0.33 = 0.31
5 5 5
I/I0

I/I0
I/I0

0 0 0
140 160 140 160 140 160
x (µm) x (µm) x (µm)

Fig. 4 Simulation of the topological control for a small waist. a Numerical simulation of the control of the final state after a propagation distance
L ¼ 2:5 mm for an initial beam waist W 0 ¼ 10 μm (I0 ¼ U PW ¼ 5:33 ´ 105 W/m2). Axis t expresses time of detection, while x is the beam transverse
0 0
coordinate. b Initial beam intensity: a super-Gaussian wave centered at x ¼ 150 μm. c–e Peregrine soliton generation: beam intensity (c) at t ¼ 12 s, and
(d) at t ¼ 64 s, during the formation of the Peregrine soliton, while (e) exhibits the Peregrine soliton profile at t ¼ 70 s. f, g Higher-order Peregrine soliton
generation: beam intensity at f t ¼ 85 s, and g t ¼ 100 s, where the Peregrine soliton is alternately destroyed and reformed

at time t  30 s. Referring to Eq. (6), we obtain the normalized illustrated in Fig. 5e–g. Each PS has two-phase signatures: a
velocity v ¼ v=v0 , with v0 ¼ L=t . longitudinal smooth phase shift of 2π and a transversal
rectangular phase shift profile, with height π and basis as wide
as the PS width47,48. Such signatures are here both experimentally
Peregrine solitons emergence. Figure 4 illustrates the numeri- demonstrated. From Fig. 5e, which shows the interference pattern
cally determined dynamics at smaller values of the beam waist during the first PS occurrence, we obtain the longitudinal phase
(W 0 ¼ 10 μm), a regime in which the generation of single PSs is shift behavior in Fig. 5g, by a cosinusoidal fitting along the central
evident. The intensity profile is reported in Fig. 4a. As shown in propagation outline. Figure 5f reports the experimental transver-
Fig. 1b, one needs to carefully choose W 0 for observing a RWs sal phase shift profile along x. A comparison with the measured
generation without the DSWs occurrence. For W 0 ¼ 10 μm, the interference fringes is also illustrated in the inset, which directly
super-Gaussian wave (Fig. 4b) generates a PS (Fig. 4c–e). The shows the phase jump (topological defect). Stressing the signi-
following dynamics shows the higher-order PS emergence ficance of these results is very important, because they are a proof
(Fig. 4f, g), each order with a higher genus. of the topological control: the genus is determined by the input
Figure 5a–g report the experimental results for the case waist and time of detection. Indeed, the longitudinal phase
W 0 ¼ 30 μm. Observations of the Peregrine-like soliton genera- shift represents the transition from genus 0 to 2, whereas the
tion are shown, both in intensity (Fig. 5a–d) and in phase transverse PS phase shift outline unveils the value g ¼ 2, equal to
(Fig. 5e–g). For a small initial waist, a localized wave, well the number of phase jumps (first from 0 to π, then again from π
described by the PS (Fig. 5b, d), forms and recurs without a to 0). This is summarized in Fig. 5h, which sketches numerical
visible wave breaking. This dynamics is in close agreement with simulations of phase behavior at W 0 ¼ 10 μm, normalized in
simulations in Fig. 4d–g, where the PS is repeatedly destroyed and ½π; π. Figure 5h gives a picture of genera changes, PS
generated, each time at a higher order. Phase measurements are occurrence and phase discontinuities. The genus is zero and the

NATURE COMMUNICATIONS | (2019)10:5090 | https://doi.org/10.1038/s41467-019-12815-0 | www.nature.com/naturecommunications 5


ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12815-0

a I/I0
Peregrine solitons
5

200
4

x (µm) 100
2

1
1 2 3

0
0
0 20 60 100 140
t (second)

b c d
4 1 4 2 5 3
3 3
I/I0

I/I0

I/I0
2 2
1 1
0 0 0
50 100 150 200 50 100 150 200 50 100 150 200
x (µm) x (µm) x (µm)

e 140
f
Phase shift (units of )

1.5

Y X
0  2 1
x (µm)

120
0.5

0
90 120 150
100 x (µm)
40 50
t (second)

g h
2 155
150
Phase shift (units of )

145

155
x (µm)

1
150

145
0
40 50 0 20 40 60 80 100 120
t (second) t (second)

Fig. 5 Experimental topological control for a small waist. a Observation of optical intensity I=I0 for an initial beam waist W 0 ¼ 30 μm. Axis t is time of
output detection, x is the transverse direction. In this regime, we observe Peregrine-soliton-like structures formation (see Fig. 1b) [the colored scale goes
from 0 (dark blue) to 5 (bright yellow)]. c, d Intensity outlines corresponding to numbered dashed lines in (a): the blue lines are experimental waveforms,
the pink continuous lines are fitting functions according to the analytical PS profile. e–h Phase measurements (e–g) and simulations (h) of the Peregrine
soliton. The detected interference pattern during the first PS generation is reported in (e), corresponding to (b). The jump from 0 to 2π along the white
dashed line corresponds to the transition from g ¼ 0 to g ¼ 2. The black dashed line highlights the jump, shown in (g). The experimental transversal
phase shift profile along x is reported in (f), showing the expected π shift corresponding to (b). Error bars represent standard deviation. The inset shows the
corresponding area of the measured interference fringes on the transverse plane. Phase simulations at W 0 ¼ 10 μm are reported in the bottom panel in
(h) [the colored scale goes from π (bright yellow) to π (dark blue) (0 is green)]. Top panel sketches Fig. 4a, for at-a-glance correspondence between
genera changes, PS occurrence and phase discontinuities

phase profile is flat until the first PS occurrences. After that, the Discussion
phase value changes and the phase transverse profile presents two The topological classification of nonlinear beam propagation by
jumps of π. the genera of the Riemann theta functions opens a new route to
The statistical properties of the PS intensity are illustrated in control the generation of extreme waves. We demonstrated the
Supplementary Information (Suppl. Fig. 2f, g), and they confirm topological control for the focusing box problem in optical pro-
the occurrence of RWs in the small box regime. pagation in photorefractive media. By using the time-dependent

6 NATURE COMMUNICATIONS | (2019)10:5090 | https://doi.org/10.1038/s41467-019-12815-0 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12815-0 ARTICLE

photorefractive nonlinearity, we could design the final state of the Experimental setup. A y-polarized optical beam at wavelength λ ¼ 532 nm from
wave evolution in a predetermined way and explore all the pos- a continuous 80 mW Nd:YAG laser source is focused by a cylindrical lens down
to a quasi-one-dimensional beam with waist U 0 ¼ 15 μm along the y-direction.
sible dynamic phases in the nonlinear propagation. The initial box shape is obtained by a mask of tunable width, placed in proximity
Such a novel control strategy enabled the first observation of of the input face of the photorefractive crystal. A sketch of the optical system
the continuous transition from dispersive shock to rogue waves is shown in Fig. 3a. The beam is launched into an optical quality specimen
and soliton gases, demonstrating that different extreme wave of 2:1ðxÞ ´ 1:9ðyÞ ´ 2:5ðzÞ mm K0:964 Li0:036 Ta0:60 Nb0:40 O3 (KLTN) with Cu and V
phenomena are deeply linked, and also that a proper tuning of impurities (n0 ¼ 2:3). The crystal exhibits a ferroelectric phase transition at the
Curie temperature T C ¼ 284 K. Nonlinear light dynamics are studied in the
their topological content in their nonlinear evolution allows paraelectric phase at T ¼ T C þ 8 K, a condition ensuring a large nonlinear
transformations from one state to another. The further numerical response and a negligible effect of small-scale disorder53. The time-dependent
and experimental analysis reported in Supplementary Informa- photorefractive response sets in when an external bias field E is applied along y
tion proves that this new control paradigm in third-order media (voltage V ¼ 500 V). To have a so-called Kerr-like (cubic) nonlinearity from the
has a broad range of validity, where it is not affected by linear photorefractive effect, the crystal is continuously pumped with an x-polarized
15 mW laser at λ ¼ 633 nm. The pump does not interact with the principal beam
effects, like modulation instability or loss, but its nature is gen- propagating along the z axis and only constitutes the saturation intensity I S . The
uinely nonlinear. spatial intensity distribution is measured at the crystal output as a function of
In conclusion, our result is the first example of the topological the exposure time t by means of a high-resolution imaging system composed of
control of integrable nonlinear waves. This new technique casts an objective lens (NA ¼ 0:5) and a CCD camera at 15 Hz.
In the present case, evolution is studied at a fixed value of z (the crystal output)
light on dispersive shock waves and rogue wave generation. It is by varying the exposure time t. In fact, the average index change grows and
general, not limited to the photorefractive media, and can be saturates according to a time dependence well defined by the saturation time
extended to other nonlinear phenomena, from classical to τ  100 s once the input beam intensity, applied voltage, and temperature have
quantum ones. These outcomes are not only important for fun- been fixed.
damental studies and control of extreme nonlinear waves, but
further developments in the use of topological concepts in non- Data availability
linear physics can allow innovative applications for engineering All data are available in this submission.
strongly nonlinear phenomena, as in spatial beam shaping for
microscopy, medicine and spectroscopy, and coherent super- Received: 24 February 2019; Accepted: 24 September 2019;
continuum light sources for telecommunication.

Methods
Photorefractive media. Starting from Maxwell’s equations in a medium with a
third-order-nonlinear polarization, in paraxial and slowly varying envelope
approximations, one can derive the propagation equation of the complex optical
References
1. Gardner, C. S., Greene, J. M., Kruskal, M. D. & Miura, R. M. Method for
field envelope Aðx; y; zÞ:
solving the Korteweg-de Vries equation. Phys. Rev. Lett. 19, 1095–1097 (1967).
1 2 k 2. Zakharov, V. E. & Shabat, A. B. Exact theory of two-dimensional self-focusing
{∂z A þ ∇ A þ δnðIÞA ¼ 0; ð7Þ and one-dimensional self-modulation of waves in nonlinear media. Sov. Phys.
2k n0
JETP 34, 62–69 (1972).
with z the longitudinal coordinate, x; y the transverse coordinates and n ¼ 3. Gardner, C. S., Greene, J. M., Kruskal, M. D. & Miura, R. M. Korteweg-de
n0 þ δnðIÞ the refractive index, weakly depending on the intensity Vries equation and generalizations. VI. methods for exact solution. Comm. on
I ¼ jAj2 ðδnðIÞ << n0 Þ. Pure and Appl. Math. 27, 97–133 (1974).
Equation (7) is the nonlinear Schrödinger equation (NLSE) and rules laser 4. Ablowitz, M. J., Kaup, D. J., Newell, A. C. & Segur, H. The inverse scattering
beam propagation in centrosymmetric Kerr media. For PR, the refractive index transform-Fourier analysis for nonlinear problems. Stud. Appl. Math. 53,
perturbation depends also parametrically on time, i.e., δn ¼ δnðI; tÞ. In fact, the 249–315 (1974).
amplitude of the nonlinear self-interaction increases, on average, with the exposure 5. Pierangeli, D. et al. Observation of Fermi-Pasta-Ulam-Tsingou recurrence and
time up to a saturation value, on a slow timescale, typically seconds for peak its exact dynamics. Phys. Rev. X 8, 041017–041025 (2018).
intensities of a few kWcm−2 51. 6. Mussot, A. et al. Fibre multi-wave mixing combs reveal the broken symmetry
In our centrosymmetric photorefractive crystal, at first approximation of Fermi-Pasta-Ulam recurrence. Nat. Photon. 12, 303–308 (2018).
δn ¼  δn02 f ðtÞ, with f ðtÞ the response function. δn0 includes the electro-optic 7. Akhmediev, N. N., Eleonskii, V. M. & Kulagin, N. E. Exact first-order
1 þ II solutions of the nonlinear Schrödinger equation. Theor. Math. Phys. 72,
S

effect coefficient49–51. For weak intensities I << I S , we obtain a Kerr-like regime 809–818 (1987).
with δn ¼ 2δn0 II f ðtÞ, apart from a constant term. We consider the case ∂y A  0 8. Hoefer, M. A. et al. Dispersive and classical shock waves in Bose-Einstein
S
(strong beam anisotropy), thus we look for solutions of the ð1 þ 1Þ-dimensional condensates and gas dynamics. Phys. Rev. A 74, 023623–023646 (2006).
NLSE for the envelope A  Aðx; zÞ: 9. Wan, W., Jia, S. & Fleischer, J. W. Dispersive superfluid-like shock waves in
nonlinear optics. Nat. Phys. 3, 46–51 (2007).
1 2 10. El, G. & Hoefer, M. Dispersive shock waves and modulation theory. Phys. D
{∂z A þ ∂ A þ 2ρðtÞjAj2 A ¼ 0; ð8Þ
2k x 333, 11–65 (2016).
11. Solli, D. R., Ropers, C., Koonath, P. & Jalali, B. Optical rogue waves. Nature
with ρðtÞ ¼ 2πδn
λI S f ðtÞ and the field envelope initial profile
0
450, 1054–1057 (2007).
 pffiffiffiffi
I 0 for jxj  12 W 0 12. Bonatto, C. et al. Deterministic optical rogue waves. Phys. Rev. Lett. 107,
Aðx; 0Þ ¼ : ð9Þ 053901–053905 (2011).
0 elsewhere
13. Dudley, J. M., Dias, F., Erkintalo, M. & Genty, G. Instabilities, breathers and
One obtains Eq. (8) from Eq. (1) through the transformation (3). We stress that, rogue waves in optics. Nat. Photon. 8, 755–764 (2014).
in this case, the dispersion parameter depends on time, as follows from Eq. (4). 14. Akhmediev, N. et al. Roadmap on optical rogue waves and extreme events. J.
Opt. 18, 063001–063036 (2016).
Numerical simulations. We solve numerically Eq. (1) by a one-parameter- 15. Ablowitz, M. & Clarkson, P.A. Solitons, Nonlinear Evolution Equations and
depending beam propagation method (BPM) with a symmetrized split-step in the Inverse Scattering (Cambridge University Press, 1991).
code core52. We use a high-order super-Gaussian initial condition 16. Kivshar, Y. S. & Agrawal, G. P. Optical Solitons (Academic Press, Elsevier
(   ) Science, 2003).
1 ξ 24 17. Bulso, N. & Conti, C. Effective dissipation and nonlocality induced by
ψðξ; ζ ¼ 0Þ ¼ q exp  : ð10Þ
2 l nonparaxiality. Phys. Rev. A 89, 023804–023810 (2014).
18. Ghofraniha, N., Conti, C., Ruocco, G. & Trillo, S. Shocks in nonlocal media.
For each temporal value, Eq. (1) solutions have different dispersion parameter ϵ Phys. Rev. Lett. 99, 043903–043906 (2007).
and final value of ζ, because from Eq. (3) it reads ζ fin ¼ ϵðtÞkW
4L
2 , where L is the 19. Marcucci, G., Braidotti, M., Gentilini, S. & Conti, C. Time asymmetric
0
crystal length. In Fig. 2 and 4, we show the numerical results. The propagation in quantum mechanics and shock waves: Exploring the irreversibility in
time considers ψðξ; ζ fin Þ, which corresponds to detections at end of the crystal. nonlinear optics. Ann. Phys. 529, 1600349–1600365 (2017).

NATURE COMMUNICATIONS | (2019)10:5090 | https://doi.org/10.1038/s41467-019-12815-0 | www.nature.com/naturecommunications 7


ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12815-0

20. El, G. A., Gammal, A., Khamis, E. G., Kraenkel, R. A. & Kamchatnov, A. M. 48. Xu, G. et al. Phase evolution of Peregrine-like breathers in optics and
Theory of optical dispersive shock waves in photorefractive media. Phys. Rev. hydrodynamics. Phys. Rev. E 99, 012207–012214 (2019).
A 76, 053813–053830 (2007). 49. DelRe, E., Ciattoni, A., Crosignani, B. & Tamburrini, M. Approach to space-
21. Ghofraniha, N., Gentilini, S., Folli, V., DelRe, E. & Conti, C. Shock waves in charge field description in photorefractive crystals. J. Opt. Soc. Am. B 15,
disordered media. Phys. Rev. Lett. 109, 243902–243905 (2012). 1469–1475 (1998).
22. Kartashov, Y. V. & Kamchatnov, A. M. Two-dimensional dispersive shock 50. DelRe, E. & Palange, E. Optical nonlinearity and existence conditions for
waves in dissipative optical media. Opt. Lett. 38, 790–792 (2013). quasi-steady-state photorefractive solitons. J. Opt. Soc. Am. B 23, 2323–2327
23. Erkintalo, M., Genty, G. & Dudley, J. M. On the statistical interpretation of (2006).
optical rogue waves. Eur. Phys. J. Sp. Top. 185, 135–144 (2010). 51. DelRe, E., Crosignani, B. & Di Porto, P. Photorefractive solitons and their
24. Erkintalo, M. et al. Higher-order modulation instability in nonlinear fiber underlying nonlocal physics. Progr. Opt. 53, 153–200 (2009).
optics. Phys. Rev. Lett. 107, 253901–253905 (2011). 52. Kutz, J. N. Data-Driven Modeling and Scientific Computing (Oxford
25. Chabchoub, A., Hoffmann, N., Onorato, M. & Akhmediev, N. Super rogue University Press, 2013).
waves: observation of a higher-order breather in water waves. Phys. Rev. X 2, 53. Pierangeli, D. et al. Super-crystals in composite ferroelectrics. Nat. Commun.
011015–011020 (2012). 7, 10674–10680 (2016).
26. Akhmediev, N., Dudley, J. M., Solli, D. R. & Turitsyn, S. K. Recent progress in
investigating optical rogue waves. J. Opt. 15, 060201–060209 (2013).
27. Agafontsev, D. S. & Zakharov, V. E. Integrable turbulence and formation of Acknowledgments
rogue waves. Nonlinearity 28, 2791–2821 (2015). We acknowledge M. Conforti, S. Gentilini, P.G. Grinevich, A. Mussot, P.M. Santini,
28. Kibler, B., Chabchoub, A., Gelash, A., Akhmediev, N. & Zakharov, V. E. S. Trillo, and V.E. Zakharov for fruitful conversations on related topics. We thank MD
Superregular breathers in optics and hydrodynamics: omnipresent modulation Deen Islam for technical support in the laboratory. The present research was supported
instability beyond simple periodicity. Phys. Rev. X 5, 041026–041037 (2015). by PRIN 2015 NEMO project (grant number 2015KEZNYM), H2020 QuantERA
29. Pierangeli, D., DiMei, F., Conti, C., Agranat, A. J. & DelRe, E. Spatial rogue QUOMPLEX (grant number 731473), H2020 PhoQus (grant number 820392), PRIN
waves in photorefractive ferroelectrics. Phys. Rev. Lett. 115, 093901–093906 2017 PELM (grant number 20177PSCKT), Sapienza Ateneo (2016 and 2017 programs),
(2015). and Ministry of Science and Technology of Taiwan (105-2628-M-007-003-MY4).
30. Toenger, S. et al. Emergent rogue wave structures and statistics in spontaneous
modulation instability. Sci. Rep. 5, 10380–10387 (2015). Author contributions
31. Suret, P. et al. Single-shot observation of optical rogue waves in integrable G.M. and D.P. equally contributed to this work. G.M., R.K.L., and C.C. conceived the
turbulence using time microscopy. Nat. Commun. 7, 13136–13143 (2016). idea and the theoretical framework; D.P., E.D., and C.C. conceived its experimental
32. Tikan, A. et al. Universality of the Peregrine soliton in the focusing dynamics realization. G.M. and C.C. developed the theoretical background. G.M. performed the
of the cubic nonlinear Schrödinger equation. Phys. Rev. Lett. 119, numerical simulations. D.P. carried out experiments and data analysis. A.J.A. designed
033901–033906 (2017). and fabricated the photorefractive crystal. All authors discussed the results and wrote
33. Gelash, A. A. & Agafontsev, D. S. Strongly interacting soliton gas and the paper.
formation of rogue waves. Phys. Rev. E 98, 042210–042221 (2018).
34. Gelash, A. et al. Bound state soliton gas dynamics underlying the noise-
induced modulational instability. arXiv:1907.07914 (2019). Competing interest
35. Zakharov, V. E., Sobolev, V. V. & Synakh, V. C. Behavior of light beams in The authors declare no competing interests.
nonlinear media. J. Exp. Theor. Phys. 33, 77–81 (1971).
36. El, G. A. & Kamchatnov, A. M. Kinetic equation for a dense soliton gas. Phys.
Rev. Lett. 95, 204101–204104 (2005).
Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/s41467-
37. Redor, I., Barthélemy, E., Michallet, H., Onorato, M. & Mordant, N.
019-12815-0.
Experimental evidence of a hydrodynamic soliton gas. Phys. Rev. Lett. 122,
214502–214507 (2019).
Correspondence and requests for materials should be addressed to G.M.
38. Wan, W., Dylov, D. V., Barsi, C. & Fleischer, J. W. Diffraction from an edge in
a self-focusing medium. Opt. Lett. 35, 2819–2021 (2010).
Peer review information Nature Communications thanks Giovanna Tissoni, Nail
39. Picozzi, A. et al. Optical wave turbulence: towards a unified nonequilibrium
Akhmediev and the other, anonymous, reviewer(s) for their contribution to the peer
thermodynamic formulation of statistical nonlinear optics. Phys. Rep. 542,
review of this work. Peer reviewer reports are available.
1–132 (2014).
40. El, G. A., Khamis, E. G. & Tovbis, A. Dam break problem for the focusing
Reprints and permission information is available at http://www.nature.com/reprints
nonlinear Schrödinger equation and the generation of rogue waves.
Nonlinearity 29, 2798–2836 (2016).
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
41. Audo, F., Kibler, B., Fatome, J. & Finot, C. Experimental observation of the
published maps and institutional affiliations.
emergence of peregrine-like events in focusing dam break flows. Opt. Lett. 43,
2864–2867 (2018).
42. Biondini, G. Riemann problems and dispersive shocks in self-focusing media.
Phys. Rev. E 98, 052220–052226 (2018). Open Access This article is licensed under a Creative Commons
43. Biondini, G. & Lottes, J. Nonlinear interactions between solitons and Attribution 4.0 International License, which permits use, sharing,
dispersive shocks in focusing media. Phys. Rev. E 99, 022215–022221 (2019). adaptation, distribution and reproduction in any medium or format, as long as you give
44. Bertola, M., El, G. A. & Tovbis, A. Rogue waves in multiphase solutions of the appropriate credit to the original author(s) and the source, provide a link to the Creative
focusing nonlinear Schrödinger equation. Proc. R. Soc. A 472, Commons license, and indicate if changes were made. The images or other third party
20160340–20160351 (2016). material in this article are included in the article’s Creative Commons license, unless
45. Xu, G., Conforti, M., Kudlinski, A., Mussot, A. & Trillo, S. Dispersive indicated otherwise in a credit line to the material. If material is not included in the
dam-break flow of a photon fluid. Phys. Rev. Lett. 118, 254101–254105 article’s Creative Commons license and your intended use is not permitted by statutory
(2017). regulation or exceeds the permitted use, you will need to obtain permission directly from
46. Kibler, B. et al. The Peregrine soliton in nonlinear fibre optics. Nat. Phys. 6, the copyright holder. To view a copy of this license, visit http://creativecommons.org/
790–795 (2010). licenses/by/4.0/.
47. Randoux, S., Suret, A., Chabchoub, P., Kibler, B. & El, G. Nonlinear spectral
analysis of Peregrine solitons observed in optics and in hydrodynamic
experiments. Phys. Rev. E 98, 022219–022230 (2018). © The Author(s) 2019

8 NATURE COMMUNICATIONS | (2019)10:5090 | https://doi.org/10.1038/s41467-019-12815-0 | www.nature.com/naturecommunications

You might also like