1 4976107

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Nonlinear optical properties of As20S80 system chalcogenide glass using Z-scan

and its strip waveguide under bandgap light using the self-phase modulation
L. E. Zou, P. P. He, B. X. Chen, and M. Iso

Citation: AIP Advances 7, 025003 (2017); doi: 10.1063/1.4976107


View online: http://dx.doi.org/10.1063/1.4976107
View Table of Contents: http://aip.scitation.org/toc/adv/7/2
Published by the American Institute of Physics

Articles you may be interested in


Fabrication of three-dimensional micro-nanofiber structures by a novel solution blow spinning device
AIP Advances 7, 025002 (2017); 10.1063/1.4973719

Atomic layer deposition and properties of mixed Ta2O5 and ZrO2 films
AIP Advances 7, 025001 (2017); 10.1063/1.4975928

Atom probe microscopy of zinc isotopic enrichment in ZnO nanorods


AIP Advances 7, 025004 (2017); 10.1063/1.4976299

Influence of substrate on structural and transport properties of LaNiO3 thin films prepared by pulsed laser
deposition
AIP Advances 7, 025005 (2017); 10.1063/1.4971842

High quality boron-doped epitaxial layers grown at 200°C from SiF4/H2/Ar gas mixtures for emitter formation
in crystalline silicon solar cells
AIP Advances 7, 025006 (2017); 10.1063/1.4976685

3D transient model to predict temperature and ablated areas during laser processing of metallic surfaces
AIP Advances 7, 025007 (2017); 10.1063/1.4976725
AIP ADVANCES 7, 025003 (2017)

Nonlinear optical properties of As20 S80 system


chalcogenide glass using Z-scan and its strip
waveguide under bandgap light using
the self-phase modulation
L. E. Zou,1,a P. P. He,1 B. X. Chen,2 and M. Iso3,b
1 Department of Physics, Nanchang University, Nanchang 330031, China
2 School of Optical-Electrical and Computer Engineering, University of Shanghai
for Science and Technology, Shanghai 200093, China
3 Department of Chemical Engineering, Tokyo University of Agriculture and Technology,

Tokyo 184-8588, Japan


(Received 10 December 2016; accepted 30 January 2017; published online 7 February 2017)

Optical nonlinearities in the undoped As20 S80 , low doped P2 As20 S78 and Sn1 As20 S79
chacogenide glasses are investigated by using Z-scan method. These experiments
show that at 1064 nm the figure of merit (FOM) for As20 S80 is ∼1.02, while
for Sn1 As20 S79 increases to ∼1.42, and for P2 As20 S78 decreases to ∼0.83. These
resulted data indicate the addition of Sn in As20 S80 system chalcogenide glass
can enhance FOM due to creating narrow energy gaps. In addition, the self-phase
modulation (SPM) width experiment for Sn1 As20 S79 strip waveguide displays that
the full width half maximum (FWHM) of spectral width increases approximately
0.8 nm under the induction of bandgap light, meaning that the bandgap light can
induce to enhance its optical nonlinearity with the nonlinear refractive index of
n2 5.27×10 14 cm2 /W. © 2017 Author(s). All article content, except where oth-
erwise noted, is licensed under a Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/). [http://dx.doi.org/10.1063/1.4976107]

I. INTRODUCTION
Chalcogenide glasses are formed by individual chacogenides of III–V group elements (S, Se,
and Te) combining with other elements such as As,Sb, Ge, Ga,1 having optically highly nonlinear (2
orders of magnitude higher than in silica) and wide transparency window of 1 µm∼12 µm from the
near-to the mid-infrared.2,3 Furthermore, for these glasses, there are very low two-photon absorption
and no free carrier absorption in telecommunication wavelengths,4,5 leading them as one of the best
materials applying into all-optical signal processing.6 The As-S and As-Se glass compositions are
well studied due to their chemical durability and the possibility for preparing fibers and waveguides.
Their photonic devices including ultrafast optical signal processing,7 supercontinuum generation,8,9
Raman amplification,10 and mid-infrared fiber laser 11 have been recently demonstrated, promoting
effectively the nonlinear photonics progress.12 Compared with As40 S60 which has stable structure and
durable chemical property due to its critical coordination number (∼2.4), As20 S80 is a less constrained
soft glass with low average coordination number (∼2.2), and has lots of lone pairs of electrons with
strong activity playing an important role in polarizability.13,14 These lone pairs of electrons are easy
to cause lattice distortion, and make the effective correlation energy of electron with negativity. In
As20 S80 system chalcogenide glass, we had investigated their photoinduced and thermal refractive
index changes.15–17 On this basis, their stripe waveguides were prepared, and an optical stopping
effect based on their waveguides was demonstrated.15,16,18

a
Electronic addresses: Linerzou@gmail.com
b
M. Iso is deceased (Oct 5, 2016)

2158-3226/2017/7(2)/025003/10 7, 025003-1 © Author(s) 2017


025003-2 Zou et al. AIP Advances 7, 025003 (2017)

In order to further understand the optical properties of As20 S80 system chalcogenide glasses, and
then improve their nonlinearity, we consider P in the third-period element with its lattice vibration
absorption edge of ∼0.1 eV higher than ∼0.08 eV for As, and Sn in IVA group with its lattice
vibration absorption edge of ∼0.06 eV less than As and S, respectively. These suitable doping elements
with appropriate concentration can change polaron concentration and energy levels in energy gap in
As20 S80 chalcogenide material, leading to the changes of motion state of these electrons and their
peripheral lattice distortion, and also affecting chalcogenide glass nonlinear optical properties. In
this paper, the As20 S80 , P2 As20 S78 and Sn1 As20 S79 films were prepared, and their nonlinear optical
characteristics were investigated by using Z-scan method. The experimental results show that the
addition of Sn can enhance the value of FOM (FOM = n2 / βλ, n2 is nonlinear refractive index, β is a
two-photon absorption coefficient), however for the addition of P, the value of FOM can descend. In
addition, in SPM experimental, the FWHM spectrum with spread of about 0.8 nm was observed in
Sn1 As20 S79 stripe waveguide induced by bandgap light, which means that the induction of bandgap
light can enhance the optical nonlinearities with n2 5.27×10 14 cm2 /W.

II. SAMPLE PREPARATION AND EXPERIMENT


As20 S80 system glasses (As20 S80 , P2 As20 S78 , and Sn1 As20 S79 ) were prepared by the sinter
method as evaporation source. These components (As, S, Sn, P) according to the mole number
of these compositions were heated to 800 ◦ C in the rocking furnace for 12 h, and then cooled to be
bulk glasses in room temperature. These glasses were grounded into powder as evaporation sources
for preparation of films. As20 S80 system chalcogenide glasses are amorphous materials, Fig. 1 is
an optical image of As20 S80 system chalcogenide glass. Fig. 1 (a) shows bulk glass in quartz tube,
and Fig. 1 (b) shows film on quartz substrate. Due to lower-doped P2 As20 S78 and Sn1 As20 S79 in
As20 S80 system chalcogenide glass, their optical images are almost same in the color difference.
These As20 S80 system films were prepared onto thin polished quartz disks (so as to work with sam-
ples with smoother surfaces) by the use of an evaporation vacuum coating technology in a vacuum
of 1×10 3 Pa with the deposition rate of about 10 Ås 1 , and their thickness was about 300 µm via
repeated the same operation. To keep the compositions of the film close to the glass compositions,
the substrate temperature was controlled to be less than 80 ◦ C. The main difference between the bulk
glass and film structure of As20 S80 is that when the bulk glass turns into the evaporation-deposited
film, some bond transitions occur. There are S8 rings in As20 S80 bulk glass with its peaks at 151 cm 1 ,
217 cm 1 and 474 cm 1 . However, As20 S80 film structures do not have S8 rings, instead, have Sn chains
with the very broad spectral range 420–520 cm 1 .19 Fig. 2 shows the Raman spectra of the As20 S80 ,
Sn1 As20 S79 and P2 As20 S78 film, manifesting that there are the range 260–420 cm-1 mainly featured
by As-S vibrations whose dominant features are 345 and 363 cm-1 corresponding to symmetric and
antisymmetric modes of AsS3 pyramidal units containing As-S bonds, and the very broad spectral

FIG. 1. An ptical image of As20 S80 system chalcogenide glass: (a) bulk glass in quartz tube; (b) film onto quartz substrate.
025003-3 Zou et al. AIP Advances 7, 025003 (2017)

FIG. 2. Raman spectra of the As20 S80 , Sn1 As20 S79 and P2 As20 S78 film.

range 420–520 cm-1 representing various types of S-S chains including 116 cm-1 for only two S atoms
in As20 S80 system films.20–22 The low addition of Sn or P to the system modifies the As-S structure,
leading a shift of the some peaks and occurrence of new band around 120–220 cm-1 .23,24 Hence the
structure of amorphous As20 S80 system film is better described as a mixture of AsS3 pyramids and
S–S chains. The refractive index of As20 S80 , Sn1 As20 S79 and P2 As20 S78 film sample measured using
a prism-coupling technique is 2.306, 2.384 and 2.288 at λ = 632.8 nm, respectively.
The investigation of nonlinear optical characteristics of As20 S80 system glass films was carried
out by Z-scan technique which is a precise method based on self-focusing to measure the nonlinear
optical response of materials, as shown in Fig. 3. The normalized transmittance in the far field as a
function of sample position Z measured with respect to the plane of the laser Gaussian beam waist
was measured, thereby the nonlinear coefficient of material was calculated.25 In Fig. 3, the sample
is placed near the focus of a converging lens, and can be moved back and forth by a micrometric
table along the beam propagation Z-axis. A variable diameter aperture was fixed at a distance of
150 cm from the lens focus, and a detector D2 was followed closely for the far field measurement.
In our case, the aperture diameter is adjusted to be about 1 mm, the closed aperture scheme in Z-
scan configurations is performed for the determination of sign and value of the nonlinear refraction
index;25 while the aperture diameter is greater than 15 mm, or the aperture is removed, the open
aperture scheme is performed for the nonlinear absorption.25 In our experiments, a Nd:YAG laser
radiation with the wavelength of 1064 nm is employed as the excitation source delivering linearly
polarized 17 ps single pulses with a 10 Hz repetition rate, and its beam waist radius ω0 is about
65 µm. The thickness of samples was about 300 µm. The distance ∆Z between maximal and minimal
transmittance of sample along Z direction is about 2∼3 cm. The radiation energy detected by D2 is
normalized in respect to the input radiation detected by D1 to avoid the influence of non-stability of
the output power of laser. In order to make sure the experimental precision, Z-scan experiment was

FIG. 3. Experimental setup for Z-scan system.


025003-4 Zou et al. AIP Advances 7, 025003 (2017)

first performed on ∼0.5 mm thick silicon sample to verify the calibration of the setup before carrying
our Z-scan experiment on chalcogenide samples. It is because the nonlinear absorption of silicon is
well documented due to its wide application in integrated photonics.

III. RESULTS AND DISCUSSION


A. The nonlinear refraction index by the closed aperture Z-scan
Fig. 4(a), (b) and (c) show the normalized transmittances T (Z) of As20 S80 , Sn1 As20 S79 and
P2 As20 S78 chalcogenide film samples in the closed aperture Z-scan configuration as functions of the
position Z with respect to focal point, respectively. The parameter ∆T PV is to easily measured by
∆T PV =T P T V (T P and T V are the normalized peak and valley transmittances), and the nonlinear
refractive index n2 (when only third order nonlinear optical processes are taken into account) can be
determined as follows.25
∆TPV
n2 = (1)
0.404(1 − S)0.25 kLeff I0
Where S = 1 − exp(−2ra2 /ωa2 ) (for an closed aperture Z-scan, S<1) is the aperture linear transmittance
with r a the aperture radius and ωa denoting beam radius at aperture in the linear regime, L eff is the
sample effective thickness by the description of Leff = [1 − exp(−αL)]/α with L the sample thickness
and α the linear absorption coefficient, and I 0 is the radiation intensity in the focal point.
These curves with a valley–peak feature in Fig. 4 exhibit that there are positive nonlinearity in
the As20 S80 system chalcogenide glass films, and then this causes the occurrence of the self-focusing
process. The nonlinear refractive index n2 (λ = 1.064 µm, I 0 =2×108 W/cm2 , α = 0.098 cm 1 ) of
As20 S80 , Sn1 As20 S79 and P2 As20 S78 chalcogenide film samples are calculated by Eq. (1). For undoped
As20 S80 , n2 is ∼4.11×10 14 cm2 /W, while for Sn1 As20 S79 increases to ∼4.68×10 14 cm2 /W. This
indicates that the addition of Sn to As20 S80 can enhance the nonlinear refractive index. However, as
n2 for P2 As20 S78 sample decreases to ∼3.89×10 14 cm2 /W, the addition of P to As20 S80 causes to
decrease the nonlinear refractive index.
B. The nonlinear absorption by the open aperture Z-scan
As we know, removal of the aperture completely eliminates the sensitivity to nonlinear refraction
index. However, the open aperture Z-scan will still be sensitive to nonlinear absorption. In this open
aperture Z-scan configuration, the main contribution to the absorption is due to the nonlinear absorp-
tion, linear absorption in sample is negligible. For the open aperture (S=1) scheme, the normalized
transmittance can be determined as25
ln(1 + q0 )
T (Z, S = 1) = (2)
q0
Where
. .
q0 = βI0 Leff (1 + Z 2 z02 ) (3)

In Eq. (3), z0 (= kω02 /2, ω0 is laser beam waist radius, and k = 2π/λ) is the beam diffraction length.
When the open aperture (S=1) Z-scan is performed, the nonlinear absorption coefficient β can be
deduced by Eq. (3).
Fig. 5(a), (b) and (c) present the normalized transmittances T (Z) of As20 S80 , Sn1 As20 S79 and
P2 As20 S78 chalcogenide glass film samples in the open aperture Z-scan, respectively. The measured
β (λ = 1.064 µm, I 0 =2×108 W/cm2 ) values of chalcogenide glass film As20 S80 , Sn1 As20 S79 and
P2 As20 S78 are ∼0.38×10 9 cm/W, ∼0.31×10 9 cm/W and ∼0.44×10 9 cm/W respectively, showing
these chalcogenide glass films have large nonlinear absorption coefficient.
The measured nonlinear parameters (the nonlinear refractive index n2 , and the nonlinear absorp-
tion coefficient β) for As20 S80 , Sn1 As20 S79 and P2 As20 S78 chalcogenide glass film samples are shown
in Table I. The nonlinear figure of merit, FOM=n2 / βλ, calculated at a wavelength of 1.064 µm for
As20 S80 system chalcogenide glass films in Table I shows that their values for FOM are between
0.80 and 1.45. In these samples, FOM for undoped As20 S80 is ∼1.02, while for Sn1 As20 S79 is the
025003-5 Zou et al. AIP Advances 7, 025003 (2017)

FIG. 4. The normalized transmittances of chalcogenide glass film samples in the closed aperture Z-scan configuration: (a)
As20 S80 , (b) Sn1 As20 S79 , and (c) P2 As20 S78 .

largest and increases to ∼1.42, and for P2 As20 S78 decreases to ∼0.83. The differences of FOM in
these chalcogenide glass film samples can be explained completely from our previous analysis about
As20 S80 system glass electron configuration.26,27 Fig. 6 shows the visible light spectrum of As20 S80 ,
Sn1 As20 S79 and P2 As20 S78 glass film samples with thickness of about 1 µm. In Fig. 6, the short-wave
absorption edge with a red shift of about 8 nm appears in Sn1 As20 S79 sample compared with undoped
025003-6 Zou et al. AIP Advances 7, 025003 (2017)

FIG. 5. The normalized transmittances of chalcogenide glass film samples in the open aperture Z-scan: (a) As20 S80 , (b)
Sn1 As20 S79 , and (c) P2 As20 S78 .

As20 S80 , meaning the addition of Sn can cause the appearance of anti-bonding state levels of coordi-
nation bond and then make energy gaps narrow. This leads to the decrease of two-photon absorption
in infrared region and then decreases nonlinear absorption coefficient. However, for P2 As20 S78 film
samples, the addition of P in As20 S80 system can increase the depth of potential well of the self-
trapped polaron states and then provide wide energy gaps, which is coincident with its the short-wave
absorption edge with a blue shift about 7.2 nm shown in Fig. 5 and is in favor of two-photon absorption
in infrared region and then increases nonlinear absorption coefficient.
025003-7 Zou et al. AIP Advances 7, 025003 (2017)

TABLE I. The nonlinear refractive index n2 , the nonlinear absorption coefficient β, and FOM(=n2 /βλ) of As20 S80 ,
Sn1 As20 S79 and P2 As20 S78 film samples.

Sample Abs.edge (µm) λ (µm) n2 ×10 14 (cm2 /W) β × 10 9 (cm/W) FOM (n2 /βλ)

As20 S80 0.440 1.064 4.11±0.2 0.38±0.02 ∼1.02


Sn1 As20 S79 0.448 1.064 4.68±0.2 0.31±0.02 ∼1.42
P2 As20 S78 0.433 1.064 3.89±0.2 0.44±0.03 ∼0.83

FIG. 6. The visible light spectrum of As20 S80 , Sn1 As20 S79 and P2 As20 S78 films.

IV. NONLINEAR PROPERTIES OF SN1 AS20 S79 STRIP WAVEGUIDE UNDER BANDGAP
LIGHT INDUCTION
Chalcogenide glasses have high non-linearity, as a promising nonlinear material for integrated
all-optical devices. Their waveguide structure with high nonlinearity is one of the most important
integrated optical components as the parametric conversion, supercontinuum generation, all-optical
signal regeneration and so on. In addition, Chalcogenide glasses have high photosensitivity to bandgap
light. In this section, we examine the nonlinear properties of our waveguides under induction of band
gap light, presenting experimental results for the measured nonlinear refractive index and spectral
broadening due to the self-phase modulation (SPM) for intense pulses propagating and the induction
of bandgap light in waveguide.
A double optical coupler SPM system was established to investigate the nonlinear optical prop-
erties of chalcogenide waveguide under the bandgap light induction, as shown in Fig. 7. A 1572 nm
input pulses with 800 fs duration from the KTP OPO laser source with the repetition rate of 81MHz
were coupled into the single-mode optical fiber using a focusing lens, was employed as a probe light,
and A 446.1 nm wavelength He-Cd laser with a power of 180 mW (its wavelength falls within the
optical band gap of As20 S80 system glass) was adopted as a pumping light (as a role of the induction
of bandgap light, meaning excitation of free electrons into the conduction band) in the same cou-
pling way. The two beams were multiplexed by WDM, and then butt-coupled into the chalcogenide
waveguide via a fiber lens. The transmitted output was monitored using an optical spectrum analyzer
(OSA).
The n2 value is estimated from the following Eq. (4).28
4
∆v 2 = ∆v02 (1 + √ γ2 P2 ) (4)
3 3
where
2πn2
γ= Leff (5)
λAeff
025003-8 Zou et al. AIP Advances 7, 025003 (2017)

FIG. 7. The setup for a double optical coupler self-phase modulation system.

∆v is the output spectral width (FWHM) for optical pulse after phase shift. ∆v0 is initial spectral
width, L eff is the effective waveguide length (Leff = [1 − exp(−αL)]/α, α is the linear absorption
coefficient, and L is the waveguide length), P is the input peak power, λ is the optical wavelength,
Aeff is the effective mode field area.
In our experiment, the addition of Sn in As20 S80 system film displays higher FOM in above sam-
ples, so Sn1 As20 S79 waveguide is selected to investigate nonlinear optical properties of waveguide. A
5 cm long Sn1 As20 S79 strip waveguide with the width of 5 µm and the depth of 1 µm was fabricated
on quartz substrates with two polished faces using lithography technology based on the photosensi-
tivity of Sn1 As20 S79 film under irradiation at ultraviolet wavelength,17 and then was employed in the
setting up of Fig. 7. The transmission loss of this Sn1 As20 S79 strip waveguide is about 0.7 dB/cm,
and the entire coupling efficiency including the end face coupling and Fresnel reflection in Fig. 6 is

FIG. 8. (a): The optical spectrum of probe light; (b): The output normalized spectrum for optical pulses propagating in
Sn1 As20 S79 strip waveguide by the self-phase modulation effect under the He-Cd laser operation (OFF and ON).
025003-9 Zou et al. AIP Advances 7, 025003 (2017)

about 10 dB. The effective area of the fundamental mode for this waveguide is computed to be about
2.32 µm2 at 1550 nm. Fig. 8(a) shows the optical spectrum of probe light (input signal). The output
normalized spectrum for optical pulses propagating in Sn1 As20 S79 strip waveguide in which the peak
power of probe light is about 40 W is shown in Fig. 8(b), indicating the spectral width is broadened
due to the SPM. The n2 without the induction of bandgap light (off He-Cd laser) is estimated to
(4.43±0.2)×10 14 cm2 /W, closed to that obtained by Z-scan method. More importantly, in Fig. 8(b)
under the induction of He-Cd laser (on He-Cd laser) the output FWHM spectral width of SPM of
the photo-inducing Sn1 As20 S79 strip waveguide is about 0.8 nm more than that without the induction
(off He-Cd laser). Then the calculated n2 value for the photo-inducing Sn1 As20 S79 strip waveguide is
(5.27±0.3)×10 14 cm2 /W, indicating the nonlinear effect is enhanced under the induction of bandgap
light (446.1 nm) in Sn1 As20 S79 strip waveguide.
It is well known that the refractive index of chalcogenide glasses film under the irradiation
increases, while it recovers immediately to a certain degree after the irradiation is switched off.29,30
This is a dynamic reversible process. In chalcogenide glass material, the nonlinear distortion of atomic
extranuclear electron orbital is prone to be produced due to photon effect under the irradiation, and
then causes a change of refractive index, resulting in the change of nonlinear phenomena. The increase
of refractive index is mainly due to the polarization of electron around sulfur atom or its center under
irradiation, enhancing the density of charges and neutral state defects in material. Moreover, it also
leads to the dynamic changes of absorption coefficient and optical bandgap, and these changes are
intrinsically linked with the nonlinear optical properties. In addition, the nonlinear refractive index
n2 is also consistent with the Miller’s empirical formula n2 ∝ n0 4 ×10 10 (esu), and increase with the
linear refractive index n0 .

V. CONCLUSION
The nonlinear optical properties in As20 S80 system chalcogenide glasses including the undoped
As20 S80 , tin doped Sn1 As20 S79 and phosphorus doped P2 As20 S78 were investigated. The nonlinear
refractive index n2 obtained by the closed aperture Z-scan experiments shows that n2 can increase to
∼4.68×10 14 cm2 /W in the tin doped Sn1 As20 S79 , while it can decrease to ∼3.89×10 14 cm2 /W in
the phosphorus doped P2 As20 S78 , compared with n2 of ∼4.11×10 14 cm2 /W in the undoped As20 S80 .
The nonlinear absorption coefficient β measured by the open aperture Z-scan experiments indicates
that these chalcogenide glass films have large nonlinear absorption coefficient of about an order of
10-10 cm/W. The FOM calculated in these chalcogenide glass film samples expresses that the value for
FOM for Sn1 As20 S79 is the largest, ∼1.42, while for P2 As20 S78 decrease to ∼0.83, compared with the
value of ∼1.02 for As20 S80 . For the photo-inducing Sn1 As20 S79 strip waveguide under the induction
of band gap light (446.1 nm), its value of n2 increases to ∼5.27×10 14 cm2 /W, which expresses that
the induction of band gap light can achieve the stronger optical nonlinear effect. These results show
the addition of Sn can increase the optical nonlinearities in As20 S80 system glass for its applying for
all-optical signal processing.

ACKNOWLEDGMENTS
This work was supported by the National Natural Science Foundation of China (No. 61465008,
and No. 60967003).
1 A. B. Seddon, “Chalcogenide glasses: A review of their preparation, properties and applications,” J. Non-Cryst. Solids 184,

44–50 (1995).
2 J. M. Harbold, F. O. Ilday, F. W. Wise, J. S. Sanghera, V. Q. Nguyen, L. B. Shaw, and I. D. Aggarwal, “Highly nonlinear
As-S-Se glasses for all-optical switching,” Opt. Lett. 27(2), 119–121 (2002).
3 Z. G. Lian, Q. Q. Li, D. Furniss, T. M. Benson, and A. B. Seddon, “Solid microstructured chalcogenide glass optical fibers

for the near-and mid-infrared spectral regions,” IEEE Photon. Technol. Lett. 21(24), 1804–1806 (2009).
4 D. Hewak, “The promise of chalcogenides,” Nat. Photonics 5, 474 (2011).
5 K. Suzuki, Y. Hamachi, and T. Baba, “Fabrication and characterization of chalcogenide glass photonic crystal waveguides,”

Opt. Express 17(25), 22393–22400 (2009).


6 A. Prasad, C. J. Zha, R. P. Wang, A. Smith, S. Madden, and B. Luther-Davies, “Properties of Ge As Se
x y 1-x-y glasses for
all-optical signal processing,” Opt. Express 16(4), 2804–2815 (2008).
025003-10 Zou et al. AIP Advances 7, 025003 (2017)

7 B. J. Eggleton, T. D. Vo, R. Pant, J. Schroder, M. D. Pelusi, D. Choi, S. J. Madden, and B. Luther-Davies, “Photonic
chip based ultrafast optical processing based on high nonlinearity dispersion engineered chalcogenide waveguides,” Laser
Photonics Rev. 6(1), 97–114 (2012).
8 Q. Zhang, M. Li, Q. Hao, D. Deng, H. Zhou, H. Zeng, L. Zhan, X. Wu, L. Liu, and L. Xu, “Fabrication and characterization

of on-chip optical nonlinear chalcogenide nanofiber devices,” Opt. Lett. 35(22), 3829–3831 (2010).
9 H. G. Dantanarayana, N. Abdel-Moneim, Z. Tang, L. Sojka, S. Sujecki, D. Furniss, A. B. Seddon, I. Kubat, O. Bang, and

T. M. Benson, “Refractive index dispersion of chalcogenide glasses for ultra-high numerical-aperture fiber for mid-infrared
supercontinuum generation,” Optical Materials Express 4(7), 1444–1455 (2014).
10 W. Shulin, X. Yinsheng, Z. Yaxun, Z. Peiqing, C. Feifei, and D. Shixun, “Gain characteristics of Tm3+ /Ho3+ -Co-doped

chalcogenide glass fibers,” J. Chin. Ceram. Soc. 41(7), 141–146 (2013).


11 J. S. Sanghera, L. B. Shaw, and I. D. Aggarwal, “Chalcogenide glass-fiber-based mid-IR sources and applications,” IEEE

J. Sel. Top. Quantum Electron 15, 114–119 (2009).


12 B. J. Eggleton, B. Luther-Davies, and K. Richardson, “Chalcogenide photonics,” Nat. Photonics 5, 141–148 (2011).
13 V. M. Lyubin and V. K. Tikhomirov, “Novel photoinduced effects in chalcogenide glasses,” J. Non-Cryst. Solids 135(1),

37–48 (1991).
14 P. K. Gupta, “Non-crystalline solids: Glasses and amorphous solids,” J. Non-Cryst. Solids 195(1), 158–164 (1996).
15 L. E. Zou, B. X. Chen, L. Chen, and Y. F. Yuan, “Fabrication of an As S stripe waveguide with an optical stopping effect
2 8
by exposure to ultraviolet irradiation,” App. Phys. Lett. 88, 153510 (2006).
16 L. E. Zou, B. X. Chen, H. S. Lin, H. Hamanaka, and M. Iso, “Fabrication and propagation characterization of As S
2 8
chalcogenide channel waveguide made by UV irradiation annealing,” App. Opt. 48(33), 6442–6447 (2009).
17 B. Sun, B. X. Chen, G. R. Sui, G. D. Wang, L. E. Zou, M. Iso, and H. Hamanaka, “Photo-induced refractive index change of

amorphous tin-doped As2 S8 films and its application to strip waveguide fabrication,” J. App. Phys. 105(9), 094501 (2009).
18 L. Zou, S. Yao, B. Chen, H. Hamanaka, and M. Iso, “Optical stopping effect on As S system film waveguide,” 2012
20 80
Conference on Lasers and Electro-Optics, CLEO 2012, San Jose, CA, USA, 2012.05.06–11, Post.
19 Z. Liner, W. Gouri, S. Yun, C. Baoxue, and M. Iso, “As S planar waveguide: Refractive index changes following an
2 8
annealing and irradiation and annealing cycle, and light propagation features,” Journal of Semiconductors 32(11), 112004
(2011).
20 F. Kyriazis and S. N. Yannopoulos, “Colossal photostructural changes in chalcogenide glasses: Athermal photoinduced

polymerization in Asx S100 x bulk glasses revealed by near-bandgap Raman scattering,” Appl. Phys. Lett. 94, 101901
(2009).
21 F. Kyriazis, A. Chrissanthopoulos, V. Dracopoulos, M. Krbal, M. Frumar, T. Wagner, and S. N. Yannopoulos, “Effect of

silver doping on the structure and phase separation of sulfur-rich As–S glasses: Raman and SEM studies,” J. Non-Cryst.
Solids 355, 2010 (2009).
22 R. Naik, S. K. Parida, C. Kumar, R. Ganesan, and K. S. Sangunni, “Optical properties change in Sb S Se thin films by
40 40 20
light induced effect,” J. Alloys Compd. 522, 172–177 (2012).
23 L. L. Cheng, M. H. Liu, M. X. Wang, S. C. Wang, G. D. Wang, Q. Y. Zhou, and Z. Q. Chen, “Preparation of SnS films using

solid sources deposited by the PECVD method with controllable film characters,” J. Alloys Compd. 545, 122–129 (2012).
24 S. A. Fayek, “The effects of Sn addition on properties and structure in Ge–Se chalcogenide glass,” Infrared Physics &

Technology 46, 193–198 (2005).


25 M. Sheik-Bahae, A. A. Said, T. H. Wei, D. J. Hagan, and E. W. Van Stryland, “Sensitive measurement of optical nonlinearities

using a single beam,” IEEE Journal of Quantum Electronics 26(4), 760–769 (1990).
26 W. Guan-De, C. Bao-Xue, W. Ping, S. Guo-Rong, L. E. Zou, H. Hamanaka, and M. Iso, “Mechanism of optical stopping

effect of arsenic sulfide amorphous waveguide,” Acta Phys. Sin (in Chinese). 60(7), 074224 (2011).
27 L. E. Zou, S. T. Yao, B. X. Chen, Y. Shen, H. Hamanaka, and M. Iso, “Recovery response of optical stopping effect on

P2 As20 S78 and Sn1 As20 S79 film waveguide,” AIP Adances 2, 012146 (2012).
28 G. P. Agrawal, Nonlinear Fiber Optics, fourth edition (Academic Press, New York, 2007).
29 L. E. Zou, I. V. Kabakova, E. C. Mägi, E. Li, C. Florea, I. D. Aggarwal, B. Shaw, J. S. Sanghera, and B. J. Eggleton,

“Efficient inscription of Bragg gratings in As2 S3 fibers using near bandgap light,” Opt. Lett. 38(19), 3850–3853 (2013).
30 I. V. Kabakova, L. Zou, G. A. Brawley, C. Florea, I. D. Aggarwal, J. S. Sanghera, E. C. Mägi, E. Li, and B. J. Eggleton,

“Dynamics of photoinduced refractive index changes in As2 S3 fibers,” App. Opt. 51(30), 7333–7338 (2012).

You might also like