Optical Materials: Sciencedirect

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Optical Materials 94 (2019) 9–14

Contents lists available at ScienceDirect

Optical Materials
journal homepage: www.elsevier.com/locate/optmat

In situ measurements of photoexpansion in As2 S3 bulk glass by atomic force T


microscopy
Thiago V. Morenoa, Vitor S. Zanutoa, Nelson G.C. Astratha, Givanildo R. Silvab,
Eduardo J.S. Fonsecab, Samuel T. Souzab, Donghui Zhaoc, Himanshu Jainc, Luis C. Malacarnea,∗
a
Departamento de Física, Universidade Estadual de Maringá, Maringá, PR, 87020-900, Brazil
b
Grupo de Óptica e Nanoscopia (GON), Instituto de Física, Universidade Federal de Alagoas, Maceió, AL, 57072-900, Brazil
c
Department of Materials Science and Engineering, Lehigh University, Bethlehem, PA, 18015, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Photoinduced expansion is a special property of chalcogenide glasses, which makes them useful for various
Chalcogenide glasses applications in photonics. A clear understanding of the kinetics and the topography of photoinduced effects is
Photoinduced effects fundamental to improve potential applications. A detailed study on bulk glass is performed using in situ Atomic
Atomic Force Microscopy Force Microscope measurements. The corresponding expected thermal contribution is obtained by numerical
simulation using Finite Elemental Analysis method. The results show a significant difference between the dy-
namics of metastable/permanent and transitory effects, with the permanent photoexpansion being much larger
in magnitude and slower in kinetics than the transient or reversible changes. The former effect is of athermal
origin. However, the dynamic response of the latter is similar to that expected from thermoelastic expansion,
suggesting a close correlation between the two effects, which remains to be established.

1. Introduction characteristics of the PD and PE are appreciably different, with these


two phenomena being almost independent of each other. The dynamics
Chalcogenide glasses exhibit a variety of photoinduced effects such of PD and PE are described by the stretched exponential function
as photoexpansion (PE), photocontraction(PC), photodarkening(PD) [21–24]
and defect creation when exposed to light with energy comparable to β
their optical bandgap. For example, PC is observed in thermally eva- Δh (t ) = A [1 − e−(t / τ ) ], (1)
porated as-deposited amorphous arsenic sulphide (a − As2 S3) thin film where τ and β are the effective response time and the dispersion
[1], whereas PE is observed in the well-annealed film of the same parameter, respectively. τ and β depend on the thickness of films and
composition but different thermal history [2]. These photoinduced ef- are different for PD and PE [22,23,25]. For PD the parameter τ increases
fects depend on different mechanisms and physical conditions, such as strongly with increasing thickness, while β decreases as the thickness
temperature, thermal history, photon energy, light intensity, light po- increases. In addition to the stretched exponential behaviour, other
larization, thickness, etc [3–7]. In a − As2 S3 films, two distinct beha- models have also been proposed to describe the dynamics of photo-
viours of PE and PD are observed, one transitory and the other per- induced effects [14,15].
manent that is metastable [8–10]. When the annealed As2 S3 films or Most of the past studies of photoinduced effects were conducted ex
glass flakes are irradiated by a sub band gap high-intensity laser, the so- situ thus providing information only about the metastable changes
called giant photoexpansion is observed [3,5]. These photoinduced [3–5,15,17,26]. In order to obtain insight into the kinetics of transitory
changes in As2 S3 films (or bulk samples) can be recovered with an- part, in situ measurements were also performed, but only in films or
nealing at the glass-transition temperature. Such characteristics of very thin bulk samples by employing optical methods
chalcogenide glasses lead to a range of possible applications [11–13]. [7–9,21–24,27,28].
The nature and the origin of the photostructural changes and their The athermal origin of photoinduced effects is assumed due to the
relation with the optical changes have been discussed in a variety of shape and the amplitude of resulting permanent deformation, which
models [13–20]. It was demonstrated that the temporal growth follows the laser beam profile, and the large difference between


Corresponding author.
E-mail address: lcmala@dfi.uem.br (L.C. Malacarne).

https://doi.org/10.1016/j.optmat.2019.05.016
Received 12 April 2019; Received in revised form 8 May 2019; Accepted 12 May 2019
Available online 24 May 2019
0925-3467/ © 2019 Elsevier B.V. All rights reserved.
T.V. Moreno, et al. Optical Materials 94 (2019) 9–14

observed expansion and expected thermal expansion [15]. Recently, the


temperature rise and the thermoelastic expansion for a model material
As2 S3 glass were calculated, and the results showed that the contribu-
tion of expected thermal expansion may not be neglected in order to
correctly describe the dynamics of PE [29]. However, the proposed
model is valid only for samples with large dimensions due to the semi-
infinite sample approximation. For a long time illumination and finite
size samples the heat-coupling with the surrounding medium and the
elastic boundary condition can have large effect on the resulting
thermal expansion, and thus a more complete model is necessary in
order to describe the thermoelastic process. Only in situ measurements
allow a comparison of the photoexpansion with the thermoelastic
contribution since the heat spreads as a function of time, dictated by the
thermal diffusion equation, and thus it would disappear when the heat
is dissipated after switching off the laser. The metastable photoexpan-
sion is a plastic deformation which persists even after laser irradiation
has been extinguished. Fig. 1. A schematic diagram of the experimental apparatus for AFM measure-
The employed optical methods to measure in situ deformation in ments.
general give only a global evolution of the induced deformation. On the
other hand, Atomic Force Microscopy has been used to investigate the scheme [35]. The cantilever oscillation amplitude was kept constant by
dynamics and reversibility of Ag nanoparticle photodissolution in the feedback loop.
chalcogenide As2 S3 glass [30,31] and the resulting surface topology in In order to make sure that the measured displacement is not caused
direct laser writing in As − S − Se chalcogenide thin film [32] and by thermally induced bending of the cantilever, we have used glass tip
a − Se bulk samples [33]. In the latter, the topology of the emerging which does not absorb light at 532 nm . The tip was held between the
surface relief indicates that the formation mechanisms are thermally microscope lens and the sample without obstructing the excitation
induced processes generated by the local heating and involve thermo- beam. It was exposed and illuminated using the lens of the microscope,
plastic deformation, mass flow induced by surface tension gradient and allowing us to view the exact region from where surface displacement
decomposition-evaporation mechanism. Recently, it was shown that the information was being collected.
AFM setup, with a tapered glass tip attached to a tuning fork probe, is During the experiments, the excitation laser beam was absorbed by
able to give real time information of laser induced surface deformation the sample resulting in a time-dependent deformation of the surface.
[34]. This local displacement causes an oscillation amplitude variation of the
In this work, we investigate the shape and the dynamics of the laser tip because the distance between tip and sample surface is reduced. The
induced effects in As2 S3 bulk glass by taking thermal and athermal dependency between oscillation amplitude and oscillation frequency on
contributions into account. We present experimental results for the the distance between tip and surface makes the electronic feedback loop
evolution of laser-induced surface deformation by in situ AFM mea- system promptly react in order to keep a constant working distance
surement using a glass tip that does not absorb light at 532 nm . This between tip and sample. The information of the surface displacement at
setup allows a direct measurement in the illuminated area. The ther- the contact point between AFM tip and sample was monitored with
moelastic contribution to the dynamics of the photoexpansion is highly time. At each measurement, the sample position was changed and,
dependent on the boundary condition, and its effect presents a fast before the laser was switched on, a mapping was performed to avoid
diffusion governed by the thermal and elastic properties of the material repeated measurements in a previously illuminated region. After the
and surrounding medium. The thermoelastic process is dictated by a set laser was turned off, and the transitory part was relaxed, a surface map
of differential equations, which do not have analytical solutions in was obtained to show the resulting permanent deformation. A bulk
many cases. Therefore, in order to estimate the real thermal contribu- sample of As2 S3 with thickness of l = 1.72mm was used in all mea-
tion for the transient part of photoexpansion, numerical simulations are surements.
performed using Finite Elemental Analysis of the thermoelastic con- The simulations were performed by using Finite Elemental Analysis
tribution, while taking into account the finite sample size and heat- method (Comsol Multiphysics Software), which provides numerical
coupling of the sample with the surrounding air. solutions to the thermoelastic equations with realistic boundary con-
ditions imposed by the experimental geometry. The heat coupling with
2. Material and method the surrounding air and fixed back face boundary condition were in-
corporated in model calculation. For short transients with time much
A schematic diagram of the experimental apparatus for AFM mea- smaller than the characteristic diffusion time, the boundary condition
surements is presented in Fig. 1. A continuous wave (cw) linearly po- does not produce significant effect on the results. However, the
larized diode-pumped solid-state (DPSS) laser of wavelength λ = 532nm boundary conditions and the heat-coupling effect are relevant for long
was used as the excitation source (Ultralasers Model MSL-FN-532-200). time transients in order to obtain correct solution.
A mechanical shutter was used to control the exposure of the excitation
laser beam on the sample. The excitation beam was focused on the
sample surface using an objective with focal distance of 17.5 mm . The 3. Results and discussion
diameter of the excitation beam on the sample surface was ∼ 15μm and
the samples were excited with power P = 1mW . The surface displace- Fig. 2 shows the resulting permanent deformation for t = 20s and
ment measurements were carried out using an AFM (Nanonics, Multi- t = 60s of illumination, and the corresponding deformation profile for
view 4000TM) in tapped-mode working in phase feedback, with a ta- the y coordinate fixed at the maximum amplitude. The continuous lines
2 2
pered glass tip attached to a tuning fork probe (Nanonics) having a tip are the fit with a Gaussian function e−2(x − x 0) / w . For both profiles, good
diameter of 10 nm and resonance frequency of 37.5 kHz . In this mode, fitting was obtained with ω = 15μm , near the excitation beam radius.
the probe oscillates close to its resonance frequency and has brief This reproduces a well-known result that the permanent deformation
contact with the sample once in every oscillation. We used a tuning fork follows the laser beam shape, demonstrating that the permanent ex-
detection system instead of the conventional laser-PSD detection pansion is not of thermal origin. The thermoelastic contribution spreads

10
T.V. Moreno, et al. Optical Materials 94 (2019) 9–14

Fig. 2. AFM maps of As2 S3 chalcogenide glass for pulses of t = 20s and t = 60s .
Fig. 4. (Open circles) Total photoexpansion and (open square) corresponding
(Open symbols) Profile of the permanent deformation for y-coordinate fixed at
permanent deformation for t = 1s, 20s and 60s obtained from the AFM maps (as
the maximum amplitude. (Continuous lines) Fit with Gaussian function.
indicated by the arrows) after the laser is turned off. The dashed line is a visual
guide.
as a function of time leading to non-Gaussian profile, which is dictated
by the thermal diffusion equation and the thermoelastic equation
[37,38].
The dynamics of the photoinduced expansion was obtained in an on-
off measurement with the tip at the center of the beam profile. Fig. 3
shows ten measurements performed in different positions of the sample.
To insure that the regions do not present permanent deformation, a
map of each region around the point was performed before the illu-
mination. The measurements show a consistent result, with only small
variation between all data. In addition, we clearly observe the fast
partial decay of the amplitude after the laser is turned off (t = 20s ).
A longer transient (t = 60 s ) and the corresponding evolution of the
permanent photoexpansion are shown in Fig. 4. The permanent ex-
pansion was obtained from the amplitude of the maps after the re-
laxation for different time of illumination. The same permanent and
transitory behaviours are demonstrated in a sequence of on (t = 1 s )
and off (t = 10 s ) transients shown in Fig. 5-a. The result clearly shows
the dynamics of the permanent and the transitory photoexpansion ef-
fects.
In order to check this on-off behaviour, we repeated the sequence of
on-off transients with the pump-probe Thermal Mirror (TM) and
Thermal Lens (TL) techniques. In these techniques, a solid-state laser
(λ = 532 nm ) in the TM00 mode is used to induce the effect and a very

Fig. 5. Sequence of on (t = 1s ) and off transients of As2 S3 chalcogenide glass:


(a) AFM, (b) TM and (c) TL techniques.

weak He-Ne laser (λ = 632.8 nm ), with a bigger radius than the ex-
citation laser, is used to probe the effect by monitoring the central re-
gion of the reflected and transmitted light. For the reflected light, the
surface expansion works as a convex mirror, making the intensity at the
central region to decrease. For the transmitted part of probe light, the
temperature gradient works as a convergente lens (positive temperature
coefficient of the refractive index), making the intensity at the central
region to increase. Both techniques are highly sensitive to the change in
the optical absorption. A test was done to ensure that no effect was
generated by the probe laser, by turning the probe laser on-off with
excitation laser off: indeed no transitory effect was observed. A detailed
description of TM and TL techniques can be found in literature [36–38].
Fig. 3. On-off AFM transients for As2 S3 chalcogenide glass illuminated by a Fig. 5-b shows the same behaviour on the TM transient as observed
laser beam (λ = 532nm ). A large expansion is observed when the laser is turned in the AFM measurement. The only difference is that, in TM technique
on (t ≤ 20s ) and a partial relaxation when the laser is turned off (t > 20 ). the deformation of the entire surface contributes to the transient, while,

11
T.V. Moreno, et al. Optical Materials 94 (2019) 9–14

β1 β2
t t
ΔhOn (t ) = A1 ⎡1 − e−( τ1 ) ⎤ + A2 ⎡1 − e−( τ2 ) ⎤
⎢ ⎥ ⎢ ⎥ (2)
⎣ ⎦ ⎣ ⎦

for the laser-on, and


β2
⎡ t−ξ ⎞ ⎤
−⎛
ΔhOff (t ) = ΔhOn (ξ ) ⎢Δu + (1 − Δu) e ⎝ τ2 ⎠
⎜ ⎟


⎢ ⎥ (3)
⎣ ⎦

for the laser-off. ξ is the time when the laser is turned off. A good
agreement is obtained with τ1 = 3.76 s , τ2 = 0.06 s , β1 = 0.67 , β2 = 0.39,
and Δu = 0.68. The result is consistent with the fast relaxation of the
transitory part with τ1 ≫ τ2 . The values of the dispersive parameters,
0 < β < 1, are consistent with the topological origin of stretched ex-
ponential for stress and structural relaxation in glasses [39]. On the
other hand, the effective reaction time τ is highly dependent on the
laser beam power and sample thickness [23,24]. Then, a direct com-
Fig. 6. Comparative TL and TM transients. (red circle) TM signal and (black
circle) TL signal. (For interpretation of the references to colour in this figure
parison with the literature is not easy since we used a bulk sample and,
legend, the reader is referred to the Web version of this article.) in this setup, the laser is focused producing an energy density higher
than in usual transmittance measurements. However, this method al-
lows a direct determination of the dependence of this parameter with
in AFM, the tip was kept at fixed position. This explains an amplitude of
the energy density. As this was not our objective, this point would be
the transitory deformation that is bigger in TM than in AFM. TM
evaluated in future work. The expected thermal contribution shown in
technique was used here only to confirm the observed AFM transient.
Fig. 7 was obtained by numerical simulation with the use of known
For the TL transient, Fig. 5-c, after the excitation laser is turned off, the
thermal and mechanical properties of As2 S3 glass [40]. In the inset of
signal returns to the same value as at t = 0. Fig. 6 shows a comparative
Fig. 7, an enlarged view of the short time data is shown. A fast ex-
on-off TM and TL transients for 20s of excitation. The TL transient signal
pansion is observed in the beginning of the transient, followed by a
returns to the same value as at t = 0 after the laser is turned off,
slow behaviour after the initial period. This initial behaviour can be
showing no permanent effect or detectable changes in the refractive
correlated to the transitory part of the photoexpansion, and it presents a
index or optical absorption for this characteristic time. The absence of
similar characteristic time to the thermal expansion, but with the am-
photodarkening could be associated to the dependence on the thickness
plitude an order of magnitude larger than the thermal contribution.
or by the short time scale of our measurements [21,23,25]. The TM
The effect of the thermal expansion was explored by taking AFM
signal shows the permanent deformation after the laser is turned off. A
transients with the tip positioned at a distance r from the central part of
clear difference is also observed in the dynamics for both transients. In
the beam. Fig. 8 shows the decreasing of the total and permanent
TL, we have only a fast characteristic thermal evolution. The TM
photoexpansion as a function of radial distance for t = 20 s of illumi-
transient presents the initial fast increase in the amplitude followed by
nation. The data were obtained by transient measurements (see inset
a slowly characteristic photoexpansion evolution. A more detailed
plot in Fig. 8) with the tip positioned at r = 0,10,20,40, and 60 μm from
study using these methods is underway.
the center of the laser beam. We see the permanent photoexpansion
Fig. 7 shows the on-off AFM transient and data fitting to stretched
following the laser beam profile, while the transitory expansion for
exponential functions,
r < 2ω is bigger than the expected thermal contribution. However,
when r > 2ω, the expansion is basically the amplitude of the thermal

Fig. 8. Total (Solid circle) and permanent (Open circle) photoexpansion as a


Fig. 7. (Open circle) On-off AFM transients for As2 S3 chalcogenide glass illu- function of radial distance. (Open triangle) Expected thermal expansion ob-
minated by a beam laser (λ = 532nm ) of radius ω ≈ 15μm and power P = 1mW ; tained by FEA. (Continous line) Normalized radial dependence of the laser
(Dashed line) Expected thermal expansion; (Solid line) Fit with stretched beam profile (ω = 15μm ). Inset plot: The time transients obtained with the tip at
functions. Inset plot: A zoom for short time (t ≤ 1s ). r = 0,10,20,40, and 60μm from the center of the laser beam.

12
T.V. Moreno, et al. Optical Materials 94 (2019) 9–14

Table 1 photoexpansions, indicating a non-thermal origin of photoexpansion.


Physical properties of the As2 S3 glass [40,41] and air used in the simulations. However, similar dynamics of the response for the transitory and the
Parameter Units Sample Air thermoelastic expansion suggests a possible relation between both ef-
fects in corroboration with previous results in laser induced surface
Density kg m−3 3200 1.18 relief using a tightly focused continuous-wave laser in chalcogenide
Specific Heat J kg −1K−1 456.07 1005 [32,33]. Although the experiments performed here are not able to de-
Thermal Conuctivity W m−1 K−1 0.167 0.026 monstrate the connection between thermal and transitory photo-
Thermal Expansion Coef. 10−6K−1 21.5 expansion, TM technique is proposed as a method to explore the pos-
Optical Absorption Coef. cm−1 35
sible relation between both effects.
Poisson's ratio 0.24
Young's modulus 109Pa 15.86
Declaration of interests

expansion. This clearly shows that the amplitudes of both permanent The authors declare that they have no known competing financial
and transitory parts are not directly connected with thermoelastic interests or personal relationships that could have appeared to influ-
changes. Note that the heat source has the Gaussian profile; however, ence the work reported in this paper.
since the diffusion process is dictated by the thermal diffusion equation,
the heat spreads as a function of time. Thus the temperature rise profile Acknowledgements
has a shape broader than the heat source. In addition, the temperature
rise profile has a sharp shape compared with the surface displacement, The authors are thankful for the financial support from the Brazilian
caused by the mechanical inertia of the surface to expand upon tem- agencies Capes and CNPq (Process 401160/2016-5).
perature changes [37].
Using characteristic physical properties of As2 S3 glass given in References
Table 1 (illumination time of t = 20 s , Gaussian laser beam of radius
ω = 15 μm , and P = 1mW ), FEA simulations give a temperature in- [1] M. Kasai, H. Nakatsui, Y. Hajimoto, Photodepression in AsS thin films, J. Appl.
Phys. 45 (1974) 3209–3210.
crease of T ≈ 80C , with corresponding expansion around 60 nm at the [2] H. Hamanaka, K. Tanaka, A. Matsuda, S. Iizima, Reversible Photo-induced volume
central region. We have used an estimated values for the optical ab- changes in Evaporated As2S3 and As4Se5Ge1 ilms, Solid State Commun. 19 (1976)
sorption coefficient Ae = 35cm−1. However, in the literature there is 499–501.
[3] H. Hisakuni, K. Tanaka, Giant photoexpansion in As2S3 glass, Appl. Phys. Lett. 65
reported value for the optical absorption one order of magnitude bigger (1994) 2925–2927.
[42]. Assuming the value for the optical absorption as Ae = 350cm−1, [4] G. Pfeiffer, M.A. Paesler, S.C. Agarwal, Reversible photodarkening of amorphous
the numerical simulation leads to the correspondent thermoelastic ex- arsenic chalcogens, J. Non-Cryst. Sol. 130 (1991) 111–143.
[5] K. Tanaka, A. Saitoh, N. Terakado, Giant photo-expansion in chalcogenide glass, J.
pansion around 90 nm at the central region. The thermoelastic expan- Optoelectron. Adv. Mater. 8 (2006) 2058–2065.
sion is similar since in both cases the sample could be assumed to be [6] A. Lőrinczi, F. Sava, I.-D. Simandan, A. Velea, M. Popescu, Photoexpansion in
optically opaque, with almost all photon energy being absorbed by the amorphous As2S3: a new explanation, J. Non-Cryst. Sol. 447 (2016) 123125.
[7] H. Asao, K. Tanaka, Polarization-dependent photoinduced mechanical deformations
sample. Note that these values are of the same order and still much
in covalent chalcogenide glasses, J. Appl. Phys. 102 (2007) 043508.
smaller than the permanent and transitory photoexpansions observed in [8] A. Ganjoo, Y. Ikeda, K. Shimakawa, In situ photoexpansion measurements of
our experiment. These values for temperature and thermoelastic ex- amorphous films: role of photocarriers, Appl. Phys. Lett. 74 (1999) 2119–2121.
[9] A. Ganjoo, K. Shimakawa, K. Kitano, E.A. Davis, Transient photodarkening in
pansion are nearly the same as those obtained by the semi-infinite
amorphous chalcogenides, J. Non-Cryst. Solids 299–302 (2002) 917923.
approximation model without air heat-coupling of the reference Zhao [10] A. Ganjoo, H. Jain, Millisecond kinetics of photoinduced changes in the optical
et al. [29]. However, illumination for long time induces discrepancy parameters of a-As2S3films, Phys. Rev. B 74 (2006) 024201.
between FEA, which provides numerical solutions to the thermoelastic [11] H. Hisakuni, K. Tanaka, Optical fabrication of microlenses in chalcogenide glasses,
Opt. Lett. 20 (1995) 958–960.
equation with realistic boundary conditions, and the semi-infinite ap- [12] B.J. Eggleton, B. Luther-Davies, K. Richardson, Chalcogenide photonics, Nat.
proximation. Despite the clear non-correlation between the amplitude Photon. 5 (2011) 141–148.
of photoexpansion and thermoelastic expansion, the transitory part of [13] K. Tanaka, Photoinduced deformations in chalcogenide glasses, in: R. Wand (Ed.),
Amorphous Chalcogenides: Advances and Applications, CRC Press, Boca Raton, FL,
photoexpansion presents a dynamic response about the same as ex- 2014, pp. 59–87.
pected from thermoelastic changes. This possible connection between [14] Y. Gueguen, J.C. Sangleboeuf, V. Keryvin, E. Lépine, Z. Yang, T. Rouxel, C. Point,
both processes was also investigated in the study of laser induced sur- B. Bureau, X.-H. Zhang, P. Lucas, Photoinduced fluidity in chalcogenide glasses at
low and high intensities: a model accounting for photon efficiency, Phys. Rev. B 82
face relief using a tightly focused continuous-wave laser in chalco- (2010) 134114.
genide glass [32,33]. The similar characteristic time for the transitory [15] M. Buisson, Y. Gueguena, R. Laniel, C. Cantoni, P. Houizot, B. Bureau,
photoexpansion and thermal expansion suggests a possible relation J.C. Sangleboeuf, P. Lucas, Mechanical model of giant photoexpansion in a chal-
cogenide glass and the role of photofluidity, Physica B 516 (2017) 85–91.
between both effects, which needs to be explored in more detail. The
[16] S.R. Elliott, A unified model for reversible photostructural effects in chalcogenide
TM technique seems to be an alternative to exploring this correlation glasses, J. Non-Cryst. Solids 81 (1986) 71–98.
since it is possible to change the ratio between excitation and probe [17] A.V. Kolobov, H. Oyanagi, K. Tanaka, K. Tanaka, Structural study of amorphous
selenium by in situ EXAFS: observation of photoinduced bond alternation, Phys.
beam radius, which allows amplifying the contribution of each part for
Rev. B 55 (1997) 726734.
the TM signal. In addition, in this technique, we are able to control the [18] K. Shimakawa, N. Yoshida, A. Ganjoo, Y. Kuzukawa, A model for the photo-
laser power and wavelength more easily than in AFM. This study is structural changes in amorphous chalcogenides, Phil. Mag. Lett. (1997) 153–158
underway. 1998.
[19] K. Tanaka, Relation between photodarkening and photoexpansion in As2S3 glass,
Phys. Status Solidi B 249 (2012) 2019–2023.
[20] C. Lu, D. Recht, C. Arnold, Generalized model for photoinduced surface structure in
4. Conclusions amorphous thin films, Phys. Rev. Lett. 111 (2013) 105503.
[21] K. Shimakawa, Y. Ikeda, Transient responses of photodarkening and photoinduced
volume change in amorphous chalcogenide films, J. Optoelectron. Adv. Mater. 8
In conclusion, the AFM setup used in this work allowed a precise in (2006) 2097–2100.
situ investigation of the topography and the dynamics of photoexpan- [22] K. Shimakawa, N. Nakagawa, T. Itoh, The origin of stretched exponential function
sion in bulk chalcogenide glass. A clear difference between the dy- in dynamic response of photodarkening in amorphous chalcogenides, Appl. Phys.
Lett. 95 (2009) 051908.
namics of metastable and transitory responses was obtained. The nu- [23] N. Nakagawa, K. Shimakawa, T. Itoh, Y. Ikeda, Dynamics of principal photoinduced
merical simulation showed that the amplitude of the thermoelastic effects in amorphous chalcogenides: in-situ simultaneous measurements of photo-
deformation is much smaller than both permanent and transitory darkening, volume changes, and defect creation, Phys. Status Solidi C 7 (2010)

13
T.V. Moreno, et al. Optical Materials 94 (2019) 9–14

857–860. [34] S.T. Souza, E.J.S. Fonseca, C. Jacinto, N.G.C. Astrath, T.P. Rodrigues,
[24] A. Ganjoo, K. Shimakawa, H. Kamiya, E.A. Davis, J. Singh, Percolative growth of L.C. Malacarne, Direct measurement of photo-induced nanoscale surface displace-
photodarkening in amorphous As2S3 films, Phys. Rev. B 62 (2000) R14601. ment in solids using atomic force microscopy, Opt. Mater. 48 (2015) 71–74.
[25] K. Tanaka, Photoexpansion in As2S3 glass, Phys. Rev. B 57 (1998) 5163–5167. [35] K. Karrai, R.D. Grober, Piezoelectric tip-sample distance control for near field op-
[26] K. Tanaka, T. Gotoh, N. Yoshida, S. Nonomura, Photothermal deflection spectro- tical microscopes, Appl. Phys. Lett. 66 (1995) 1842–1844.
scopy of chalcogenide glasses, J. Appl. Phys. 91 (2002) 125–128. [36] R.D. Snook, R.D. Lowe, Thermal lens spectrometry - a review, Analyst 120 (1995)
[27] Y. Ikeda, K. Shimakawa, Real-time in situ measurements of photoinduced volume 2051–2068.
changes in chalcogenide glasses, J. Non-Cryst. Solids 338–340 (2004) 539–542. [37] N.G.C. Astrath, L.C. Malacarne, P.R.B. Pedreira, A.C. Bento, M.L. Baesso, Time-re-
[28] J. Ren, T. Wagner, J. Orava, T. Kohoutek, B. Frumarova, M. Frumar, G. Yang, solved thermal mirror for nanoscale surface displacement detection in low ab-
G. Chen, D. Zhao, A. Ganjoo, H. Jain, In-situ measurement of reversible photo- sorbing solids, Appl. Phys. Lett. 91 (2007) 191908.
darkening in ion-conducting chalcohalide glass, Optic Express 16 (2008) [38] N.G.C. Astrath, L.C. Malacarne, V.S. Zanuto, M.P. Belancon, R.S. Mendes,
1466–1474. M.L. Baesso, C. Jacinto, Finite-size effect on the surface deformation thermal mirror
[29] D. Zhao, H. Jain, L.C. Malacarne, P.R.B. Pedreira, Role of photothermal effect in method, J. Opt. Soc. Am. B 28 (2011) 17351739.
photoexpansion of chalcogenide glasses, Phys. Status Solidi B 250 (2013) 983987. [39] M. Potuzak, R.C. Welch, J.C. Mauro, Topological origin of stretched exponential
[30] V.P. Palumbo, A. Kovalskiy, H. Jain, B.D. Huey, Direct investigation of silver relaxation in glass, J. Chem. Phys. 35 (2011) 214502.
photodissolution dynamics and reversibility in arsenic trisulphide thin films by [40] AMIR-6 (As2S3) Data Sheet from Amorphous Materials Inc. http://www.
atomic force microscopy, Nanotechnology 24 (2013) 125706. amorphousmaterials.com/app/download/6552919404/AMTIR-6RInformation.pdf
[31] P. Khan, Y. Xu, W. Leon, K.V. Adarsh, D. Vezenov, I. Biaggio, H. Jain, Kinetics of (accessed 02 April 2019).
photo-dissolution within Ag/As2S3heterostructure, J. Non-Cryst. Solids 500 (2018) [41] M.S. Iovu, S.D. Shutov, A.M. Andriesh, E.I. Kamitsos, C.P.E. Varsamis, D. Furniss,
468–474. A.B. Seddon, M. Popescu, Spectroscopic studies of bulk As2S3 glasses and amour-
[32] I. Voynarovych, S. Schroeter, R. Poehlmann, M. Vlcek, Surface corrugating direct phous films doped with Dy, Sm and Mn, J. Optoelectron. Adv. Mater. 3 (2001)
laser writing of microstructures in ternary chalcogenide films using a continuous- 443–454.
wave super-bandgap laser, J. Phys. D Appl. Phys. 48 (2015) 265106. [42] E.A. Davis, N.F. Mott, Conduction in non-crystalline systems V. Conductivity, op-
[33] R. Bohdan, S. Molnar, I. Csarnovics, M. Veres, A. Csik, S. Kokenyesi, Optical re- tical absorption and photoconductivity in amorphous semiconductors, Phil. Mag. 22
cording of surface relief on amorphous selenium, J. Non-Cryst. Solids 408 (2015) (1970) 903–922.
57–61.

14

You might also like