Random Lasing Control with Optical Spatial Solitons in Nematic
Liquid Crystals
Armando Piccardi
1
, Sreekanth Perumbilavil
2
, Martti Kauranen
2
, Giuseppe Strangi
3
and
Gaetano Assanto
1
1
NooEL - Nonlinear Optics and OptoElectronics Lab, University “Roma Tre”, Rome, Italy
2
Laboratory of Photonics, Tampere University of Technology, Tampere, Finland
3
Physics Dept., Case Western Reserve University, Cleveland, Ohio, U.S.A.
Keywords: Random Laser, Nematic Liquid Crystals, Optical Spatial Solitons.
Abstract: We discuss the synergy of reorientational self-focusing and random lasing in a dye- doped nematic liquid
crystalline material. The laser emission resulting from amplification and multiple scattering inside the
medium can be either modulated or triggered depending on the energy of the visible pump beam and the
power of the near-infrared spatial soliton, respectively exciting the two nonlinear responses. Moreover, the
presence of the self-induced waveguide improves the properties of the emitted beam, i. e., directionality and
profile. Finally, the laser light can be re-directed by steering the spatial soliton with the aid of an external
low-frequency electric field.
1 INTRODUCTION
Random lasing occurs when multiple recurrent
scattering inside an optically active medium
provides the necessary feedback for the stimulated
emission to reach gain higher than losses (Wiersma,
2008). Since its prediction and the first experimental
evidence (Letokhov, 1967; Lawandy, 1994), random
lasers have attracted great attention, letting
researchers envision the realization of low-cost
tunable coherent light sources. As a drawback, their
emission is randomic in direction, and the resulting
profile has poor quality with respect to standard
sources (Cao, 2003). During the last decades, a
number of materials have been employed to generate
and control random laser emission, including
powders (Leonetti, 2011), biological tissues (Polson,
2004), conjugated polymers (Tulek, 2010),
semiconductor polycrystalline films (Cao, 1998),
perovskites (Safdar, 2018), and nematic liquid
crystals (Strangi, 2006). The latter provide light
scattering due to the thermal oscillations of their
weakly linked molecules. At the same time, the low
binding forces allow reorientation of the anisotropic
molecules by optical beams, allowing nonlinear self-
focusing and the generation of optical spatial
solitons (Peccianti, 2003).
In this work, we combined the reorientational
nonlinearity with the strongly scattering behaviour
of a nematic liquid crystal mixture doped with a dye
- acting as the active medium -, generating a
nematicon and random laser emission at the same
time (Perumbilavil, 2016), and studied their mutual
interaction. The paper is organized as follows: first,
we illustrate the principle of both random laser
emission and propagation of spatial solitons in
nematic liquid crystals. Then we expose the
experimental results, while the last part is dedicated
to the discussion of the results and the conclusion.
2 RANDOM LASING AND
NEMATICONS IN DYE-DOPED
NEMATIC LIQUID CRYSTALS
Nematic Liquid Crystals (NLC) are anisotropic
materials featuring high reorientational nonlinearity
owing to the torque induced rotation of the
elongated molecules due to electric fields at either
low or optical frequencies (De Gennes, 1993). Non
uniform molecular reorientation generated by finite
size beams compensates for diffraction, allowing for
self-focusing and the formation of spatial solitons
(Stegeman, 1999), i. e., optical wavepackets with
Piccardi, A., Perumbilavil, S., Kauranen, M., Strangi, G. and Assanto, G.
Random Lasing Control with Optical Spatial Solitons in Nematic Liquid Crystals.
DOI: 10.5220/0007575102890293
In Proceedings of the 7th International Conference on Photonics, Optics and Laser Technology (PHOTOPTICS 2019), pages 289-293
ISBN: 978-989-758-364-3
Copyright
c
2019 by SCITEPRESS Science and Technology Publications, Lda. All rights reserved
289
invariant profile in propagation (Segev, 1992;
Peccianti, 2000; Leo, 2004; Rotschild, 2006)
The nonlocal character of the response prevents
catastrophic collapse - typical of local media -,
supporting the stable propagation of 2D spatial
solitons, operating as self-induced waveguides for
optical signals at different wavelengths (Conti,
2003). The generation and control of spatial solitons
in nematic liquid crystals, namely nematicons, have
been demonstrated in a number of configurations,
together with their possible applications in devices
for signal addressing and processing (Piccardi,
2010; Piccardi, 2016; Izdebskaya, 2017; Laudyn,
2018) .
Figure 1: Geometry of the NLC sample where both the
visible pump and the self-confined near-infrared beams
are indicated: (a) top and (b) side view of the sample. The
blue ellipses represent the NLC molecules. (c)
Experimental set-up. P: polarizer, S: sample, OBJ:
microscope objective, BS: beam splitter, SM:
spectrometer. Acquired images of the emitted radiation
without (d) and with (e) a 6mW nematicon. The near-
infrared radiation has been filtered out.
Due to their unique optical and mechanical
properties, NLC can provide various kinds of
nonlinear behaviors, including random laser
emission when samples are doped with a fluorescent
dye (Ferjani, 2006). In fact, the strong scattering
provided by thermal molecular oscillations
combined with dye fluorescence can provide
spontaneous and eventually stimulated emission
when properly pumped, allowing amplification and
random lasing (Bolis, 2016; Perumbilavil, 2016;
Perumbilavil, 2018-1; Perumbilavil, 2018-2).
Moreover, this class of materials demonstrated
to be highly versatile, since the emission can be
easily controlled by exploiting temperature
variations or applied voltages (Wiersma, 2001; Lee,
2011), as the properties depend on molecular
distribution.
3 EXPERIMENTAL RESULTS
The set-up employed for the experiments is sketched
in Fig. 1. The sample, sketched in fig.1(a)-(b) - top
and side view, respectively - is a 100m thick cell
whose top and bottom surfaces have been rubbed in
order to obtain a uniform molecular reorientation at
45° with respect to the z axis, thus maximizing the
reorientational response. The sample is filled with a
mixture of E7 (n
||
=1.71 n
=1.52 the refractive
indices at =1064nm for extraordinary and ordinary
polarization, respectively) doped with 3%wt of
Pyrromethene 597 dye. As visible in fig. 1(c), a near
infrared cw beam at =1064nm (out of the
absorption band of the dye) is injected into the
sample with wave vector along z, a waist of about
3m and polarization along y - extraordinary
polarization -, and it generates a nematicon. A
second beam from a pulsed source at =532nm
(close to the absorption peak of the dye), with pulse
duration of 6ns and repetition rate 20Hz, is focused
with wave vector collinear with the near infrared
beam, and yields random lasing.
Figure 2: Acquired single shot spectra of output intensity -
arbitrary units, a.u. - (a) below (E=0.40J) and (b) above
(E=48J) threshold for various nematicon powers.
Preliminary measurements showed that the ordinary
pump polarization (along x) maximizes the emitted
PHOTOPTICS 2019 - 7th International Conference on Photonics, Optics and Laser Technology
290
radiation.
The evolution of the beams is observed by
collecting the light scattered out of the propagation
plane (yz) with a CCD camera, while the output
profile is recorded by a second camera imaging the
output with a microscope objective. Finally, a
spectrometer detects the output spectrum with a
resolution <1nm.
We excited the random laser with the green
beam and varied the near infrared power to
investigate the role of the self-induced waveguide on
the laser light. Fig. 1(d) and (e) show the
propagation of the green beam alone and when co-
propagating with a near infrared beam of power P
close to 6mW, respectively. As it can be seen, the
emission is confined within the nematicon
waveguide. Fig. 2 shows single shot spectra from the
output of the sample at various pump energies and
nematicon powers.
Figure 3: Acquired spectra averaged over 100 ms (a)
below (E=40J) and (b) above (E=48J) threshold. (c) In-
out characteristic versus pump energy at several
nematicon powers.
When the visible beam is at low energy (around
0.40J) the nematicon has negligible effects
irrespective of its power. Conversely, when the
energy overcomes the threshold value (0.48J) the
effect of the near infrared beam is to enhance the
emission, with the occurrence of a number of lasing
peaks of random wavelength and amplitude.
Figure 4: Single shot acquisition slightly below threshold:
without nematicon (black line) the system does not lase.
With a 6mW nematicon (red line) the system switches- to
the random lasing regime.
To rid of the stochastic character of the lasing
emission we gated the acquisition over 100ms (200
acquired spectra), obtaining the results shown in Fig.
3. The enhancement due to the nematicon is still
apparent, and the in/out characteristic (Fig. 3(c))
shows the typical lasing features, with a threshold
and a slope efficiency increasing with nematicon
power.
As shown in Fig. 4, when the pump energy is
close to (but under) the lasing threshold, the
nematicon favours the transition to the lasing
regime, demonstrating the possibility of switching
on the laser emission by optical means. Thus, the
nematicon does not only guide the emitted photons,
but enhances the lasing process, improving its
efficiency when increasing its near-infrared power
and the corresponding guided-wave confinement.
Being an effective waveguide for the emitted
photons, the nematicon also affects the transverse
profile of the emission. We collected both the
emitted radiation backscattered at the input and the
output signal. Fig. 5 compares the two typical
profiles: the speckled beam of the backscattered
radiation (Fig. 5(a)) and the smoother beam after
confined propagation within the nematicon (Fig.
5(b)).
Finally, the laser emission can also be re-
addressed by steering the nematicon waveguide. We
apply a voltage across the NLC sample, i. e. across
x. In this way the electric torque lifts the molecules
out of the propagation plane yz and the
nematicon walk-off chenges according to the voltage
Random Lasing Control with Optical Spatial Solitons in Nematic Liquid Crystals
291
Figure 5: Acquired photographs of the laser beam profiles
taken from (a) backscattered emission at the input and (b)
forward emission after propagation within the nematicon.
dependent orientation of the optic axis. Remarkably,
the laser emission follows the power flow of the near
infrared beam. We stress that the observable quantity is
the apparent walk-off, i. e., the projection of the actual
walk-off on the propagation plane. Fig. 6 shows both the
output laser profiles without and with applied voltage and
the plot of the resulting apparent walk-off versus voltage,
with a simple numerical fit.
Figure 6: Voltage-driven steering of random laser.
Photographs of output laser profiles for (a) V= 0V and V=
2V. (c) Acquired spectra averaged over 100ms as a
function of applied voltage. (d) Measured (black circles)
and calculated (red line) apparent walk-off versus applied
voltage.
4 CONCLUSIONS
We presented an innovative approach to random
lasing control in nematic liquid crystals, exploiting
the self-induced waveguide and the corresponding
molecular distribution of a spatial soliton to improve
the emission properties in terms of profile and
directionality, demonstrating also the possibility to
electrically control the emission direction.
The effects of the nematicon are not only to
guide the emitted photons, but also to enhance the
conversion efficiency. The corresponding guided-
wave random laser profile results smoother, and the
electric control of the nematicon walk-off allows
controlling the direction of the laser emitted
photons.
We believe this opens new perspectives on
application-oriented random lasing, introducing a
low cost source with electro-optic control.
Some open questions still request deeper
investigation: the actual role of the nematicon-
induced refractive index profile on the random laser
emission and on its bell-shape profile must be
addressed; the degree of coherence of random lasing
modes has to be verified, as well as their wavelength
dependence; moreover, the correlation between
spectral and spatial components of the random laser
emission must be investigated. Future studies will
thus tackle a model to account for the interaction
between the two nonlinearities, addressing the
modulation of the guest-host parameters (doping
percentage, sample geometry, etc.) for the
optimization of the laser. Other strategies for
direction control could also be implemented, in
either two- or three-dimensional geometries.
ACKNOWLEDGEMENT
Academy of Finland, FiDiPro grant no. 282858.
REFERENCES
Alberucci, A., Piccardi, A., Kravets, N., Buchnev, O.,
Assanto, G., 2015. Soliton enhancement of
spontaneous symmetry breaking. In Optica, 2(9).
Bolis, S., Virgili, T., Rajendran, S., Beeckman, J.,
Kockaert, P., 2016. Nematicon-driven injection of
amplified spontaneous emission into an optical fiber.
In Optics Letters, 41, 2245-2248.
Cao, H., et al., 1998. Ultraviolet lasing in resonators
formed by scattering in semiconductor polycrystalline
films. In Applied Physics Letters, 73, 3656.
Cao, H., 2003. Lasing in random media. In Waves Random
Media, 13,3.
Conti, C., Peccianti, M., Assanto, G., 2003. Route to
nonlocality and observation of accessible solitons.
Physics Review Letters, 91(7), 073901.
De Gennes, P. G., Prost, J., 1993. The physics of liquid
crystals. Oxford University Press.
Ferjani, S., et al., 2006. Thermal behavior of random
lasing in dye doped nematic liquid crystals. In Applied
Physics Letters, 89, 121109.
Izdebskaya, Y., Shvedov, V., Assanto, G., Krolikowski,
W., 2017. Magnetic routing of light-induced
waveguides. In Nature Communications, 8,14452.
PHOTOPTICS 2019 - 7th International Conference on Photonics, Optics and Laser Technology
292
Laudyn, U., Piccardi, A., Kwasny, M, Karpierz, M. A.,
Assanto, G., 2018, Thermo-optic soliton routing in
nematic liquid crystals. In Optics Letters, 43, 10.
Lawandy, N. M., Balachandran, R. M., Gomes, A. S. L.,
Sauvain, E., 1994. Laser action in strongly scattering
media. In Nature, 368, 436.
Lee, C.R., et al., 2011. Electrically controllable liquid
crystal random lasers below the Freedericksz transition
threshold. In Optics Express, 18(3), 2406-2412.
Leo, G; Amoroso, A., Colace, L., Assanto, G., Roussev,
R.V., Fejer, M. M., 2004. Low-threshold spatial
solitons in reverse-proton-exchanged periodically
poled lithium niobate waveguides. In Optics Letters,
29, 15.
Leonetti, M., Conti, C., Lopez, C., 2011. The mode-
locking transition of random lasers, In Nature
Photonics, 5, 615617.
Letokhov, V., 1967. Generation of light by a scattering
medium with negative resonance absorption. In
Journal of experimental and Theoretical Physics,
26(4), 835.
Peccianti, M., De Rossi, A., Assanto, G., De Luca A.,
Umeton, C., Khoo I.C., 2000. Electrically assisted
self-confinement and waveguiding in planar nematic
liquid crystal cells. In Applied Physics Letters 77, 7-9.
Peccianti, M., Conti, C., Assanto, G., De Luca, A.
Umeton, C., 2003. Routing of highly anisotropic
spatial solitons and modulational instability in liquid
crystals. In Nature, 432, 733-737.
Perumbilavil, S., Piccardi, A., Buchnev. O., Kauranen, M.,
Strangi, G., Assanto, G., 2016. Soliton-assisted
random lasing in optically pumped liquid crystals. In
Applied Physics Letters, 102, 203903.
Perumbilavil, S., Piccardi, A., Barboza, R., Buchnev, O.,
Kauranen, M., Strangi, G., Assanto, G., 2018.
Beaming random lasers with soliton control. In Nature
Communications, 9, 3863.
Perumbilavil, S., Piccardi, A., Barboza, R., Oleksandr
Buchnev, O., Kauranen, M., Strangi, G., Assanto, G.,
Molding random lasers in nematic liquid crystals. In
Optical Material Express, 8, 3864-3878.
Piccardi, A., Alberucci, A., Bortolozzo, U., Residori S.,
Assanto G., 2010. Soliton gating and switching in
liquid crystal light valve. In Applied Physics Letters,
96, 071104.
Piccardi, A., Kravets, N., Alberucci, A., Buchnev, O.,
Assanto, G., 2016. Voltage-driven beam bistability in
a reorientational uniaxial dielectric. In APL Photonics,
1, 011302.
Polson, R. C., Varden, Z. V, 2004. Random lasing in
human tissues. In Applied Physycs Letter, 85, 1289.
Rotschild, C., Alfassi, B., Cohen, O., Segev M., 2006.
Long-range interactions between optical solitons. In
Nature Physics, 2, 769774.
Safdar, A., Wang, Y., Krauss, T. F., 2018. Random lasing
in uniform perovskite thin films. In Optics Express,
26, 22.
Segev, M., Crosignani, B., Yariv, A., Fisher, B., 1992.
Spatial solitons in photorefractive media. In Physics
Review Letters, 68, 923-926.
Stegeman, G. I., Segev, M., 1999. Optical spatial solitons
and their interactions: universality and diversity. In
Science, 19, 286.
Strangi, G., Ferjani, S., Barna, V., De Luca, A., Versace,
C., Scaramuzza, N., Bartolino, R., 2006. Random
lasing and weak localization of light in dye-doped
nematic liquid crystals. In Optics Express, 14, 7737.
Tulek, A., Polson, R. C., Vardeny, Z. V., 2010. Naturally
occurring resonators in random lasing of conjugated
polymer films. In Nature Physics, 6, 303310.
Wiersma, D. S., 2008. The physics and applications of
random lasers. In Nature Physics, 4, 359.
Wiersma, D. S., Cavalieri, S., 2001. Light emission: A
temperature-tunable random laser. In Nature, 414,
708.
Random Lasing Control with Optical Spatial Solitons in Nematic Liquid Crystals
293