You are on page 1of 8

Journal of Non-Crystalline Solids 402 (2014) 236–243

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids


journal homepage: www.elsevier.com/ locate/ jnoncrysol

Coordination disordering in near-stoichiometric arsenic sulfide glass


O. Shpotyuk a,b,⁎, S. Kozyukhin c, Ya. Shpotyuk d, P. Demchenko d, V. Mitsa e, M. Veres f
a
Lviv Scientific Research Institute of Materials of SRC “Carat”, 202 Stryjska str., Lviv 79031, Ukraine
b
Institute of Physics, Jan Dlugosz University in Czestochowa, 13/15 al. Armii Krajowej, Czestochowa 42200, Poland
c
N.S. Kurnakov Institute of General and Inorganic Chemistry, 31 Leninsky Pr., Moscow 199991, Russia
d
Ivan Franko National University of Lviv, 50 Dragomanova str., 79005 Lviv, Ukraine
e
Uzhgorod National University, 32 Voloshina str., Uzhgorod 88000, Ukraine
f
Wigner Research Centre for Physics, Hungarian Academy of Sciences, H-1121 Budapest, Hungary

a r t i c l e i n f o a b s t r a c t

Article history: Compositionally-dependent deviations in structural metastability of near-stoichiometric glassy arsenic sulfides
Received 3 May 2014 As38S62, As40S60 and As42S58 associated with interbalanced chemical, medium-range and coordination
Received in revised form 18 June 2014 disordering are studied with optical spectrophotometry, micro-Raman scattering and X-ray diffraction measure-
Available online xxxx
ments related to. The glasses quenched from high enough temperatures above boiling point reveal characteristic
chemical disordering with additional amount of molecular products having “wrong” homopolar bonds, this effect
Keywords:
Chalcogenides;
being inhibited in over-stoichiometric As42S58 glass. Optical transmission measurements in the bandgap cut-off
Glasses; region demonstrate increased light scattering losses in high-temperature-quenched As38S62 and As40S60 glasses,
Non-crystalline materials; while no changes are detected in As42S58 glass. The most essential influence expressed in obvious increase in the
X-ray diffraction; effective periodicity and correlation length of the first sharp diffraction peak is registered in g-As40S60 prepared
Optical properties by slow cooling from relatively low temperatures such as 500 °C. Coordination disordering due to double-
bond-based quasi-tetrahedral S3/2As_S units, identified in respect to increased Raman-active stretching
modes near 530–540 cm−1, along with extra homopolar As\As bonds enhances a compositional misbalance
in this stoichiometric specimen, tending towards As-rich glass-forming network approaching g-As42S58. Possible
topological configurations of these structural anomalies are examined using ab-initio quantum chemical calcula-
tions with RHF/6-311G⁎ basis set, testifying in a favor of more distinct spacing arrangement of quasi-tetrahedral
units and S-compensating homopolar As\As bonds.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction justifying their functioning in different spheres. Typically, these disor-


dered media can be prepared by conventional melt-quenching route,
For a long time, the chalcogenide glasses (ChG), e.g. melt-quenched when some features of structural disordering proper to a melt occur to
vitreous alloys of chalcogen atoms (S, Se, Te) with some elements from be frozen in a glass [1,2].
IV and V groups of the Periodic table [1,2], have been in a sphere of tight Despite the lack of long-range interatomic correlations like in
interests because of a number of important applications [3–6]. Due to crystals, the effect of preparation conditions still remains essential
excellent transparency in mid-IR region, the ChG are especially attrac- for ChG in view of different levels of chemical (redistribution of co-
tive for modern optoelectronics and photonics, so that emergence of a valent bonds composing glassy network) and medium-range struc-
new field of materials research called Chalcogenide Photonics has been tural disordering (topological interlinking between glass-forming
declared recently [7]. units at more prolonged length scales stretching over a few nm) [8,9].
The ChG of binary As\S system possessing a good network-forming The medium-range disordering can be eliminated along with frozen me-
ability with full saturation of covalent bonding in a wide compositional chanical stresses in a glassy state due to annealing near glass transition
range including stoichiometric glassy g-As40S60 serve as model objects temperature Tg, while chemical disordering cannot be fully removed in
ChG due to this procedure [8]. So even stoichiometric g-As40S60
quenched from high enough temperatures Tq above boiling point Tb
⁎ Corresponding author at: Lviv Scientific Research Institute of Materials of SRC “Carat”, demonstrates broken chemical ordering, which results in a few percents
202 Stryjska str., Lviv 79031, Ukraine. Tel.: +380 032 263 83 03; fax: +380 032 294 97 35.
of “wrong” homopolar As\As and S\S bonds in addition to heteropolar
E-mail addresses: shpotyuk@novas.lviv.ua, olehshpotyuk@yahoo.com (O. Shpotyuk),
sergkoz@igic.ras.ru (S. Kozyukhin), demchenko@franko.lviv.ua (P. Demchenko),
As\S ones (such glass synthesis technological route can be conditionally
v.mitsa@gmail.com (V. Mitsa), vm@szfki.hu (M. Veres). referred to as high-temperature HT-regime). At the same time, the effect

http://dx.doi.org/10.1016/j.jnoncrysol.2014.06.013
0022-3093/© 2014 Elsevier B.V. All rights reserved.
O. Shpotyuk et al. / Journal of Non-Crystalline Solids 402 (2014) 236–243 237

of slow cooling reveals itself as very long physical aging in rapidly stretched amorphous halos registered on X-ray diffraction (XRD)
quenched ChG, producing measurable blue shift in the fundamental op- patterns (D8 Advance BRUKER, Cu Kα-radiation, Germany). Before
tical absorption edge [8,10]. Therefore, to avoid the influence of all experiments, the ChG samples were cut on plane-parallel disks
“wrong” homopolar bonding in g-As\S, the process of glass syn- of 1 mm in thickness and polished to high optical quality.
thesis should be arranged in a principally other way, exploring Atomic densities of the prepared glasses were controlled at room
slower cooling rates from lower temperatures T q, which are more temperature using the Archimedes method with distilled water, the ac-
than melting temperature T m but far below boiling point T b (such curacy being not worse than ±0.005 g/cm3.
glass synthesis route can be conditionally referred to as low- UV/VIS/NIR PerkinElmer LAMBDA 950 spectrophotometer (0.5–
temperature LT-regime). 1.4 μm) was used to measure optical transmission T of the samples
In respect to the phase diagram of binary As\S system [1,11,12], the (± 0.5%) in the fundamental absorption edge region.
homogeneous single-phase region of melt corresponding to As40S60 is Microstructure studies were performed with Raman scattering spec-
stretched down from Tb = 709 °C (determined at 1 atm.) to Tm = troscopy using Renishaw system 1000 micro-Raman spectrometer
310 °C (a more accurate value of the melting point for hydrothermally (with a typical 1–2 cm−1 resolution) equipped with charge coupling
grown As2S3 crystals gives Tm = 327 ± 2 °C [13]). Since S possesses camera. All measurements were performed at room temperature in
low Tm = 113 °C and Tb = 445 °C [14], the As (despite its higher melting back-scattering geometry. The diode laser beam of 785 nm (1.58 eV) fo-
point Tm ≈ 615 °C) can react with S even at such relatively low tem- cused by means of 50× objectives was used for scattering excitation, the
peratures (450–500 °C). So it seems principally possible to prepare excitation time being ~ 60 s. To avoid photostructural changes in the
the g-As40S60 regardless of the temperature from which the melt is studied ChG, the output power was limited by optical filters to the
cooled down within 310 b Tq b 709 °C range. Nonetheless, this reaction level of 80 μW.
is time-exhaust because of limited S solubility in solid As (in respect to Structural effects ascribed to medium-range disordering in the
[15,16], only 1.16 at.% of S can be dissolved in solid As at 500 °C). Hence, glasses under consideration were studied with powder XRD tech-
it is expected that under insufficient melt homogenization at these nique in a measuring mode of a so-called first sharp diffraction
relatively low Tq (e.g. in the LT-regime), the S can be non-uniformly peak (FSDP). This isolated peak in a reciprocal space is caused by
distributed in a glassy network resulting in specific type of structural some quasi-periodicities R (defined by the FSDP position 2Θ) in
disordering. real-space interatomic correlations occurred over length D (defined
So the melt-quenching route to prepare near-stoichiometric g- by the FSDP halfwidth) [20]. An explicit nature of the FSDP has still
As\S can be arranged employing respectively low temperatures, remained controversial, despite extensive efforts to give unambigu-
such as 450–525 °C [17–19]. The prepared ChG are expected to be ous interpretation of this phenomenon in last years. The FSDP be-
free of chemical disordering (as compared with HT-quenched haves unlikely to other XRD peaks, showing anomalous decrease
glassy alloys), but it remains unclear how structural metastability and shift towards higher scattering vectors Q with pressure and in-
will be equilibrated with medium-range disordering frozen during crease with temperature (while other XRD peaks demonstrate an
this LT-regime. To our knowledge, this specificity is often ignored opposite tendency in full respect to the Debye–Waller rule)
despite some evident controversies between ChG prepared in LT- [21–24]. In a more generalized approach, the FSDP is attributed to
and HT-regimes. This concerns, in the first hand, the uniformity in some structural entities forming medium-range ordering in a glass
network connectivity of glassy backbone, since even some room- (stretched through few Å), such as two-dimensional layers
temperature stable ESR centers (not proper to g-As\S, which typical- [21–24], atomic clusters [25] or voids [20,26].
ly demonstrate a diamagnetic response [1]) were detected in S-rich g- The FSDP-related XRD patterns in the ChG were registered in
As30S70 and g-As33S67 [19]. transmission geometry using STOE STADI P (STOE & Cie GmbH,
In this work, we try to disclose the nature of structural metastability Germany) diffractometer with linear position-sensitive detector
in g-As\S caused by insufficient melt homogenization at relatively (Cu Kα1-radiation, X-rays, curved Ge(111) monochromator on pri-
low temperatures such as 500 °C. Two batches of near-stoichiometric mary beam), as it was described in [27,28]. The ChG samples of
g-As40S60 allowing both S-rich and As-rich compositional deviations each composition synthesized in the LT- and HT-regimes were first-
will be compared for technological synthesis route arranged in the LT- ly powdered. Further, they were used to prepare special probes
and HT- regimes. sandwiched between amorphous films made of acetate polymer
having negligible absorption and no any XRD peaks in the region
2. Experimental of interest. All such probes were prepared so to keep full sample-
to-sample similarity between separate XRD measurements, the
ChG samples of stoichiometric g-As 40 S 60 , S-rich g-As 38 S 62 and probe thickness being close to one of standards (Si, Al 2 O 3 ) used
As-rich g-As 42 S 58 were synthesized in evacuated (2 · 10 − 5 Pa, for equipment calibration. Nearly the same pressure of gas mixture
ILMVAC CDK-180, Germany) and sealed fused quartz ampoules (90% Ar + 10% CH4) was maintained in the measuring chamber of
(8 mm in diameter) by conventional melt-quenching technique in position-sensitive detector through whole duration of the experi-
a rocking furnace as was described elsewhere [17–19]. The elemen- ment. The obtained XRD data were treated for profile fitting of the
tal constituents additionally purified with distillation (no worse peaks by WinPLOTR program [29]. To exclude apparatus influences
than three nines) in a total amount of 6 g were melted at maximal in different measuring cycles (pressure instabilities in a measuring
temperature of 500 °C (the LT-regime) or 850 °C (the HT-regime). camera, deviations in probe thickness, etc.), the XRD patterns
Whole synthesis procedure (lasted nearly 24 h) was carried out in were normalized in respect to the maximal peak. Both the angular
three steps: (1) the initial heating from room temperature to position 2Θ and full width at half maximum (FWHM) were chosen
200 °C with keeping at this temperature for 2 h, then (2) heating as principal parameters to describe the FSDP, their numerical values
to 450 °C followed by 2 h keeping at this temperature in a rocking in a reciprocal space (scattering vector Q and width ΔQ) being
regime, and finally, (3) heating ampoules to the above maximal parameterized as:
temperature followed by 6-hour keeping at this temperature with
only 2-hour rocking. Finally, the HT-ingots were air-quenched (g-
Q ¼ ð4π=λÞ  sinΘ; ð1Þ
As40S60-HT, g-As38S62-HT, g-As42S58-HT), and LT-ingots were slow-
ly cooled in a furnace-off regime (g-As 40 S 60 -LT, g-As 38 S 62 -LT, g-
As42S58-LT) to produce a glassy state, which was controlled visually
by characteristic conchoidal fracture and confirmed owing to wide- ΔQ ¼ ð4π=λÞ  sinðFWHM=2Þ: ð2Þ
238 O. Shpotyuk et al. / Journal of Non-Crystalline Solids 402 (2014) 236–243

In addition, the effective FSDP-related periodicity R and correlation 70


length L over which these periodicities occur [30] were estimated
using the Ehrenfest equation [31] simplified as: 60

50
As38S62-HT

T, %
R ¼ 2π=Q; ð3Þ
40 As38S62-LT

30
L ¼ 2π=ΔQ : ð4Þ
20
Atomic clusters identified as potential subjects of structural 10
disordering in binary As\S system and their possible structural config- a
urations were examined utilizing the PC-aided simulation procedure 0
600 800 1000 1200 1400
known as CINCA (the cation-interlinking network cluster approach)
[32,33]. The ab-initio quantum chemical modeling was performed Wavelength, nm
using HyperChem Release 7.5 program package based on restricted
Hartree–Fock (RHF) self-consistent field method using split-valence
70
double-zeta basis set with single polarization function 6-311G* [34,35].
The final geometrical optimization and single-point energy calculations 60
for selected molecular precursors of network-forming clusters were per-
formed employing the Fletcher–Reeves conjugate gradient method until 50
As40S60-HT
the root-mean-square gradient of 0.1 kcal/(Ǻ·mol) was reached. The 40

T, %
As40S60-LT
obtained energy of clusters was corrected on added H atoms according
to the procedure well developed elsewhere [36–38]. In such a way, the 30
full energy for each self-closed molecular cluster was obtained. Since
20
these clusters were artificially terminated by H atoms, the half energy
of S atoms bonded to H, energy of all H atoms themselves and corre- 10
sponding S\H bonds was subtracted from this full energy to distinguish
b
the total energy of “pure” network-forming cluster. Then, the corre- 0
600 800 1000 1200 1400
sponding cluster formation energy Ef was calculated as difference be-
tween this cluster energy and overall energy of constituting atoms. Wavelength, nm

3. Results and discussion


70

As it follows from numerous previous research [8,9,39–41], g-As\S 60


synthesized in the HT-regime essentially differ by content of frozen
50 As42S58-HT
“wrong” homopolar covalent chemical bonds, this kind of structural
T, %

disordering (referred to as chemical disordering, in terms of K. Tanaka 40 As42S58-LT


[8]) being simply governed by following disproportionality reaction pre-
sented in terms of main glass-forming structural units: 30

20
2AsS3=2 ↔As2 S4=2 þ S2=2 ; ð5Þ
10
c
which denotes that some part of a glassy matrix composed of heteropolar
0
As\S bonds within AsS3/2 pyramids is transformed to partially- or fully- 600 800 1000 1200 1400
polymerized As-rich (mostly As4S4 molecules involving two interlinked
Wavelength, nm
As2S4/2 units) and S-rich entities (homochain and homocyclic Sn mole-
cules with most probable n attaining 6, 8, 9, 10, 12, 18 and 20 integers
Fig. 1. Optical transmission spectra of g-As38S62 (a), g-As40S60 (b) and g-As42S58
[42,43]). This specificity of heteropolar-to-homopolar bond switching can (c) prepared in HT- and LT-regimes.
be conveniently expressed by modified disproportionality reaction given
in terms of principal covalent bonds (denoted by long lines) and their
nearest surrounding: constituent atoms (S and/or As) are not distributed uniformly in a
glassy backbone, as it can be expected for ChG prepared in the LT-
2  S2=2 As–S–AsS2=2 ↔S2=2 As–AsS2=2 þ S–S: ð6Þ regime.
The high degree of chemical disordering in the HT-quenched ChG
Chemical order broken in respect to above reaction (6) modifies the owing to molecular-like products based on “wrong” homopolar bonds
electronic structure of g-As\S due to the defect states appearing near creates some structural inhomogeneities, affecting light scattering,
band edges, so decreasing the optical bandgap [8]. these changes being expected in addition to long-wave shift in optical
At high enough temperature from which the quenching began Tq, transmission cut-off produced by reaction (6). Optical transmission
the equilibrium in reaction (6) is right-hand shifted, reflecting the ap- spectra on Fig. 1 clearly confirm this expectation for stoichiometric
pearance of large content of “wrong” homopolar bonds frozen in a g-As40S60 and S-rich g-As38S62. These ChG prepared in the HT-regime
glass. But at lower Tq, that is not greater than boiling point Tb (for g- demonstrate an obvious failure in optical transmittance at the funda-
As40S60, this value is close to 709 °C [11,12]), the appearance of mental absorption edge, enhancing towards lower photon wavelengths
molecular-like products based on homopolar chemical bonds are as compared with the same for LT-cooled samples. The Rayleigh scatter-
blocked by the energetic preference of heteropolar bonding. However, ing losses seem to be most favorable to explain such behavior [44,45].
this network-forming tendency will be disturbed, provided that Indeed, a practically unchanged saturation is observed for the near-IR
O. Shpotyuk et al. / Journal of Non-Crystalline Solids 402 (2014) 236–243 239

Table 1
Atomic densities ρ and the FSDP parameters of the studied As\S glasses prepared in HT- and LT-regimes.

Sample ρ, g/cm3 2Θ, ° FWHM, ° Q, Å−1 ΔQ, Å−1 R, Å L, Å

g-As38S62-HT 3.110 (5) 17.457 (14) 5.0 (3) 1.24 0.36 5.08 17.62
g-As38S62-LT 3.040 (5) 17.364 (12) 4.5 (3) 1.23 0.32 5.10 19.49
g-As40S60-HT 3.170 (5) 17.477 (12) 5.1 (3) 1.24 0.36 5.07 17.33
g-As40S60-LT 3.140 (5) 16.757 (11) 4.1 (3) 1.19 0.29 5.29 21.58
g-As42S58-HT 3.190 (5) 16.712 (12) 4.1 (4) 1.19 0.29 5.30 21.36
g-As42S58-LT 3.160 (5) 16.803 (11) 4.0 (3) 1.19 0.29 5.27 21.98

transmittance of these samples (up to 1400 nm), which is a concomi- while two weaker halos are observed near ~ 6° and ~34°. There is also
tant of wavelength dependence character for the Rayleigh scattering a smaller satellite peak located at the right from the FSDP (~ 21.3°),
[45]. However, no difference was detected for g-As42S58 prepared in this peak being attributed to different S allotropes present in the sam-
the HT- and LT-regimes (see Fig. 1c), testifying in a favor of negligible ples (α-cyclo-S9, cyclo-S20, α-cyclo-S8) [14,27,42,43,48,49]. In respect
effect of chemical disordering on these As-rich samples. to theoretically calculated XRD patterns shown in Fig. 3, there were no
It is noteworthy that ChG of both HT- and LT-batches clearly demon- character peaks near scattering angels proper to arsenolite As2O3
strate regular long-wave shift in spectral position of cut-off with As con- (with a most intensive line near 2Θ ≈ 14°), as it was characterized for
tent in good accordance with known concentration dependence of γ-irradiated g-As40S60 [27,28]. So parasitic oxidation processes are
optical bandgap in g-As\S system [46]. The saturation level of optical quite well inhibited in all studied glasses.
transmission is close to 70% whichever ChG composition and preparation As it follows from Fig. 2, the FSDP position does not depend on the
regime, corresponding roughly to the refractive index n approaching 2.5, highest temperature of the melt for g-As38S62 and g-As42S58, demonstrat-
which is also in good respect to the previously measured value [1,2]. ing respectively strict decrease in 2Θ from 17.5° to 16.7°, respectively.
Despite the spectral position of optical cut-off for g-As40S60 and The peak halfwidth (FWHM) drops slightly in the g-As 38 S62 -LT as
g-As38S62 prepared in the HT- and LT-regimes seeming to occur to compared with HT-quenched sample, but it remains nearly the
be interconnected through the blue shift (see Fig. 1a and b), which same in g-As42S58-HT. More essential changes occur in the FSDP pa-
is a characteristic feature of physically aged ChG [10], such a compar- rameters of stoichiometric g-As40S60. The HT-quenched sample of
ison is simply inadmissible. Indeed, the physically aged glasses being this composition g-As40 S60 -HT behaves similarly to g-As 38S 62 -HT
affected by atomic shrinkage due to aging typically demonstrate an giving 2Θ = 17.5° and FWHM = 5.1°, while the LT-quenched g-
increased density ρ (see, for instance [47]), while the ρ data from As38S62-LT is very close to both g-As 42 S58 showing typical for this
Table 1 show evidently an opposite tendency. The densities ρ of all LT- glass 2Θ = 16.8° and FWHM = 4.1°. These effects are accompanied
samples, which in this case can serve as aged ones, are obviously reduced by corresponding changes in the scattering vector Q and width ΔQ,
as compared with HT-samples (which should be expected as non-aged as well as effective periodicity R and correlation length L gathered
ChG). Hence this specificity is caused by chemical disordering owing to in Table 1.
difference in molecular-network content of the corresponding glasses The observed stability in the FSDP parameters for g-As42S58 testifies
prepared in the HT- and LT-regimes rather than by their physical aging. in a favor of unchanged medium-range disordering of this specimen
Thus, the chemical disordering due to HT-regime does not affect sig- prepared in different melt-quenching/cooling regimes. In other words,
nificantly the over-stoichiometric g-As42S58, clearly modifying optical the inner structural correlations, which determine the FSDP origin in
losses in g-As40S60 and g-As38S62 (Fig. 1), this process being not simply As-rich g-As42S58, are the same whichever medium-range and/or chem-
transferred to physical aging like in [10]. Nevertheless, it leaves prob- ical disordering frozen from a melt. This result permits to suppose that
lematic to which extend these changes are attributed (if any) to the FSDP-related medium-range correlations in this glass prepared in
medium-range disordering in these ChG. To answer this question let's the LT-regime are inter-balanced with ones occurred with respect to re-
refer to the FSDP-related correlations in the XRD patterns depicted in action (6) in the HT-regime, or alternatively, both types of structural
Fig. 2 and parameterized in Table 1. disordering are ineffective to parameterize the FSDP. The similar
All XRD patterns of the studied glasses contain three amorphous
halos in the investigated 2Θ region as it follows from Fig. 2. The FSDP
corresponds to the most intensive peak revealed at 2Θ ≈ 16.5–17.5°,

Fig. 3. Experimental XRD patterns of g-As40S60 prepared in HT- (black curve) and LT-
regimes (blue curve) as compared with theoretically calculated ones for crystalline
arsenolite As2O3 (red curve) and sulfur S α-cyclo-S9 and cyclo-S20 polymorphs (green
Fig. 2. XRD patterns of g-As38S62 (1,2), g-As40S60 (3,4) and g-As42S58 (5,6) prepared in HT- and purple curves, respectively). (For interpretation of the references to color in this figure
(1,3,5) and LT-regimes (2,4,6). Y offset is 12 mm for better presentation of the differences. legend, the reader is referred to the web version of this article.)
240 O. Shpotyuk et al. / Journal of Non-Crystalline Solids 402 (2014) 236–243

changes are characterized for S-rich g-As38S62, since no essential shift in Thus, in full agreement with these predictions, we reasonably expect
the FSDP position was detected. However, slight drop in the halfwidth of from our Raman scattering spectra shown in Fig. 4 that the LT-
this peak leading to ~10% rise in the correlation length L (from 17.62 Å synthesized ChG are influenced by abnormal coordination for some
for g-As38S62-HT to 19.49 Å for g-As38S62-LT) can be attributed to the ex- amount of As atoms in QT configurations based on double As_S
pansion in real-space region of structural correlations responsible for bonds, this anomaly being accompanied by the appearance of additional
the FSDP. homopolar As\As bonds. So despite obvious preference of direct
In contrast, the stoichiometric g-As40S60-LT as compared with HT- corner-shared linking between AsS3/2 pyramids [32,57], some amount
modified g-As40S60-HT demonstrates an obvious increase in both effec- of S3/2As_S units along with “wrong” As\As bonds have been formed
tive periodicity R and correlation length L corresponding to the FSDP in the LT-cooled glasses, being a source of structural disordering
(Table 1). Thus, the XRD patterns of g-As40S60-LT become very similar owing to insufficient melt homogenization, which in this respect can
to ones registered for LT- and HT-modifications of g-As42S58. So it can be identified as coordination disordering. Accepting a strong difference
be speculated that intrinsic decomposition in this glass is caused mainly in Raman cross sections σ for stretching As_S vibrations in S3/2As_S
by demixing of some S atoms from glassy network under these prep- units (σ = 21.9) and symmetric As\S vibrations in AsS3/2 pyramids
aration conditions in the LT-regime, ensuring an effective As-rich (σ = 34.0) as it is assumed from the first-principle calculations [50],
glass-forming network for atomic remainder (very similar to one in g- the content of such structural anomalies in g-As2S3 estimated roughly
As42S58). This effect agrees well with the observed long-wave shift of from surface areas under corresponding peaks of Raman spectra on
fundamental optical absorption edge in g-As40S60-LT approaching one
character for both g-As42S58 samples (Fig. 1b and c).
Thus, special kind of structural disordering seems to appear in the As38S62-HT x10

490
340
LT-cooled ChG due to insufficient melt homogenization at relatively As38S62-LT

530-540
470
365
low temperatures (500 °C). Deficient S distribution in a glassy network

310
stabilized at the LT-regime ensures looser glass structure with signifi- a

Intensity, a.u.
cantly reduced atomic densities ρ as it follows from the data gathered
in Table 1. It is noteworthy that this feature does not affect optical trans- 450 500 550
mission giving a sharp steepness in optical transmission edge of all LT-

165
180

530-540
prepared glasses (Fig. 1), but obviously enhances chemical misbalance

230

490
shifting ChG towards more As-enriched compositions.

470
265
To shed more light on the possible nature of this phenomenon, we
performed a comparative microstructure study of the glasses prepared
in the LT- and HT-regimes using micro-Raman scattering technique,
150 200 250 300 350 400 450 500 550 600
the normalized spectra being shown in Fig. 4. -1
In respect to these data, the glasses synthesized in the HT- and LT- Raman shift, cm
regimes differ gradually by the content of homopolar bonds. Indeed,
apart from sharp lines corresponding to Raman-active modes of AsS3/2 As40S60-HT x20
340
530-540
pyramids, which are principal glass-forming units in these ChG (the As40S60-LT

490
365

470
umbrella mode at 165 cm− 1 and dominant broad band of stretch
310

symmetric-asymmetric modes at ~ 340 cm− 1 modified by shoulders b


Intensity, a.u.

at ~ 310 and ~ 380 cm−1 [50–52]), the modes assigned to homopolar


bonding can be easily identified. In part, the intense bands at ~ 180, 450 500 550 600
230 and 365 cm− 1 can be attributed to As\As-based units such as
165

220
180

As4S4 molecules [50–54], while slight bands at 220 and 470 cm− 1
530-540
230
210

along with 490 cm−1 are caused by S\S vibrations in ring- and chain-
490
470

like fragments, respectively [50–52,55,56]. Surprisingly, all these


Raman-active modes (see Fig. 4) corresponding to homopolar bonding
(especially, As\As bonds) are more prominent in g-As40S60-LT.
150 200 250 300 350 400 450 500 550 600
From stoichiometry point, it seems that additional channel of S ap-
-1
pearance should be employed in the LT-cooled ChG to equilibrate with Raman shift, cm
these As\As bond extraction. This channel becomes evident from
Raman spectra inspection in 500–550 cm−1 range (as it follows from in- As42S58-HT x20
340

sets to Fig. 4). It is seen that the LT-cooled ChG demonstrate additional
As42S58-LT
365

wide band at ~530–540 cm−1, absent in the HT-quenched glasses. To 530-540


490
470
310

our knowledge, there were no Raman-active modes attributed to own


c
Intensity, a.u.

structural fragments of g-As40S60 in this spectral region, despite careful


inspection of whole g-As\S system previously [50–52,54,56].
180

220

To identify an eventual origin of this anomaly, let's refer to possible 450 500 550
165

theoretical predictions for Raman-active modes in g-As\S, which are


230
210

530-540

anticipated for this spectral region. Thus, in respect to the first-


490
470

principle cluster-modeling calculations based on density-functional


theory [50], the quasi-tetrahedral (QT) S3/2As_S unit can be evidenced
by the mode of double-bond As_S stretch vibrations at 537 cm− 1
(while other umbrella, symmetric-stretch and asymmetric-stretch 150 200 250 300 350 400 450 500 550 600
modes of this QT unit are expected to be neat 146, 335 and 365 cm−1, Raman shift, cm
-1

respectively, contributing mainly to modes of normal AsS3/2 pyramids).


This finding is also concomitant with the earlier ab-initio Hartree–Fock Fig. 4. Micro-Raman spectra of g-As38S62 (a), g-As40S60 (b) and g-As42S58 (c) synthesized
method calculations for different structural configurations in g-As2S3, in HT- and LT-regimes (the inset represents enhanced domain corresponding to symmet-
giving As_S group frequencies in 525–530 cm−1 domain [54]. ric stretching mode of QT S3/2As_S unit).
O. Shpotyuk et al. / Journal of Non-Crystalline Solids 402 (2014) 236–243 241

Fig. 4 can be expected at the level of a few percents. So this anomaly can
really be presented only as a specific kind of structural defects due to
non-uniform S distribution in a glassy matrix, resulting from LT-
deviations in conventional glass-preparation technology.
Coordination disordering gives an adequate explanation for the
above FSDP-related correlations in the XRD patterns of the studied
ChG (see Fig. 2), as it can be found within known Gaskel's concept on
crystalline-like ordering in melt-quenched network glasses [58]. In
Wright's extension [59], this concept justifies the model explaining
the FSDP as arising from a periodicity in the distribution of some “pseu-
do-planes” for Bragg diffraction, which represent “walls” separating a
succession of randomly-packed network cages in a glass. Flattened
Fig. 6. Ball-stick representation showing geometrically-optimized configurations of
cage structure produces inner equivalent planes at opposite sides of double-bond-based S_As2S2H4 (a) and S_As3S3.5H4 (b) clusters (red, yellow and gray
these cages, giving necessary contribution to the reciprocal space Fouri- colors are used to show As, S and H atoms, respectively). (For interpretation of the refer-
er component generating the FSDP. The ball-stick diagram of the well ences to color in this figure legend, the reader is referred to the web version of this article.)
known Zacharaisen's random network for binary glassy compounds
evolving two- and three-coordinated atoms on Fig. 5 [60] gives general The preferred one of these structural solutions can be justified with
insight on such successive cage “walls” acting as “pseudo-planes” for the ab-initio quantum chemical simulation using CINCA [32,33]. Let's con-
FSDP [59]. Within this diagram, the decrease in the scattering vector Q sider two possible configurations of network-forming clusters based
corresponding to the FSDP position with the increase in the pnictogen on QT units, e.g. As2S3 and As3S4.5, both being terminated by four H
concentration [61] can be viewed as deviations in distances between atoms to form self-closed molecular S_As2S2H4 and S_As3S3.5H4 pro-
cage “walls” due to enlarged pnictogen–pnictogen or shortened chalco- totypes suitable for calculations. The geometrically optimized configu-
gen–chalcogen bonds in a network of mostly heteropolar pnictogen– rations of these clusters are shown in the ball-stick representation in
chalcogen bonds. The appearance of fourfold-coordinated As atoms ter- Fig. 6 and corresponding parameters are gathered in Table 2. Calculated
minated by double As_S bonds essentially modifies the remaining ring distances for single As\S and double As_S bonds occur to be only
structure due to “compensating” As\As bonds pushed in the nearest slightly scattered in these clusters, being grouped roughly around 2.24
atomic environment. As a result, the flattened cage-like network be- and 2.07 Å, respectively, while As\As bonds are shorten in S_As2S2H4
comes enriched on As\As bonds, resulting in the effect similar to the molecules (2.41 Å against 2.47 Å). Bond angles on fourfold-coordinated
shift towards As-rich glass compositions. Therefore, the FSDP of stoichio- As atoms are smaller from character tetrahedral angle, consisting near
metric g-As40S60 prepared in the LT-regime, which contains maximal 106° and 107° in As2S3 and As3S4.5 clusters, respectively. Bond angle
content of QT S3/2As_S units, occurs to be parameterized like As-rich on S atom linking homopolar As\As bonds to QT unit in As3S4.5 cluster
g-As42S58 (see Table 1). is close to 100°. Thus, it should be concluded that both clusters reveal
The S-compensating homopolar As\As bonds can be immediately character elements of tetrahedral environment with only slight distor-
attached to fourfold-coordinated As atoms, being included directly in tion in As2S3 cluster having As\As bond just in its nearest surrounding.
QT unit as _As\[As(S2/2)]_S fragment of As2S3 network-forming clus- In cluster formation energy Ef calculations, this As2S3 cluster (with
ter, or, alternatively, they can be pushed away from this place in a cage- molecular S_As2S2H4 prototype) yields the second As3S4.5 one (with
like glassy network to create _As\[As(S3/2)]\[As(S2/2)]_S fragment molecular S_As3S3.5H4 prototype), which has more distant homopolar
of As3S4.5 network-forming cluster. The latter is necessary to form uni- As\As bond, the corresponding energetic balance being rationed
form network topology of ChG enriched on extra As\As bonds. as − 76.48 kcal/mol and − 77.42 kcal/mol. Such great difference tes-
tifies that S-compensating homopolar As\As bonds are rather inserted
deeply in a cage-like glassy network than being directly attached to
As_S double-bond. By accepting remarkable difference in the bond
lengths within the g-As\S system (2.28 Å for As\S and 2.49 Å for
As\As bond [1]), the enhancement in the FSDP-related periodicity R
and correlation length L is anticipated as it follows from random net-
work diagram on Fig. 5, in full respect to the experimentally observed
effect (see Table 1).
The possibility of double-bond linking in glassy arsenic chalcogen-
ides has still remained a subject of hot controversies in scientific litera-
ture as a key giving an unambiguous explanation for some anomalies
observed in compositional dependencies of their physical–chemical
properties. Thus, numerous attempts have been made to explain the
under-margin values observed in the reversing heat flow of DSC pat-
terns in As\S/Se ChG near the mean coordination number (the number
of covalent chemical bonds per atom) Z = 2.286 due to this kind of in-
teratomic bonding, thus proving compositional boundaries for so-called
optimally-constrained intermediate phase in their elastic proper-
ties [50]. Indeed, by assuming covalent bond as Lagrangian constraint
nc, two types of elementary structural units having (in per-atom deter-
mination) the number of these constraints nc approaching the space di-
mensionality (in case of 3D space nc = 3.00) can be expected for binary
g-As\S [50,51]. The first is AsS3/2 pyramids linked through common S
Fig. 5. Ball-stick representation showing pseudo-planes for “Bragg” diffraction formed by atom (corner-sharing S2/2As\S\AsS2/2 pyramidal units), which corre-
successive cage “walls” within known Zachariassen's random network diagram [26] con-
tributing to the FSDP in Wright's presentation [24] (red and yellow colors are stand for As
sponds to stoichiometric As2S3 composition (Z = 2.40), while the sec-
and S atoms, respectively). (For interpretation of the references to color in this figure leg- ond is S3/2As_S quasi-tetrahedrons, which are proper to more S-rich
end, the reader is referred to the web version of this article.) ChG (Z = 2.286). As glassy backbone had been composed only from
242 O. Shpotyuk et al. / Journal of Non-Crystalline Solids 402 (2014) 236–243

Table 2
Geometric parameters of optimized network-forming clusters in binary As\S system based on double-covalent bond and their formation energies Ef (kcal/mol).

Network (molecular) cluster Bond distance, 10−4 nm Bond angle, deg Ef

As\S As\As As_S ∠S\As\S ∠S\As_S ∠As\As_S ∠S\As\As ∠As\S\As

As2S3 2244 2463 2069 97.7 115.0 111.1 98.0 – −76.48


(S_As2S2H4) 2249 2347 101.8 114.8 97.8
2226
average 2240 2405 2069 99.8 114.9 111.1 97.9 –
As3S4.5 2247 2470 2063 100.1 116.7 – 105.4 99.7 −77.42
(S_As3S3.5H4) 2255 101.6 115.3 95.0
2245 103.0 114.7 100.7
2287 102.6 95.1
2217 102.6
2217
2217
Average 2241 2470 2063 102.0 115.6 – 99.1 99.7

these units, all corresponding ChG are expected to be optimally- while S\S vibrations in ring- and chain-like fragments (470 cm−1 and
constrained with nc = 3.00, giving a reason for so-called reversibility 490 cm−1, respectively) obviously grow. It means that concentration
window in this range of compositions. However, this is not a case of ho- excess of S in g-As\S system does not facilitate a global network con-
mogenized glassy networks prepared in conventional melt-quenching/ nectivity resulting in nc = 3.00, so that coordination disordering is rath-
cooling conditions, neither As\S [18] nor As\Se [62], since reliable er compositionally near-stoichiometric effect.
experimental evidences for double-bond-based QT structural units
S(Se)3/2As_S(Se) have not been obtained before. It means that 4. Conclusions
double-covalent bonding does not govern primary network-forming
tendencies in glassy arsenic chalcogenides, in contrast to phosphorous Structural metastability in near-stoichiometric g-As\S is composi-
chalcogenides, where such possibilities can be evidently realized [1,63]. tionally dependent in view of interbalanced chemical, medium-range
This is the first time when this type of structural anomaly (coordi- and coordination disordering, the latter being caused by insufficient
nation disordering) was found in g-As40S60 as being produced by melt homogenization at relatively low temperatures such as 500 °C.
inhomogeneities in S distribution originating from technological conditions Molecular products based on “wrong” homopolar covalent chemical
of glass preparation. Surprisingly, that double-bond linking was not found bonds in HT-prepared glasses appeared in respect to heteropolar-to-
in ChG synthesized at high over-melting temperatures, but, otherwise, homopolar bond redistribution revealing the essential effect on optical
obtained at relatively low temperatures but with insufficient melt ho- transmission cut-off due to long-wave shift and increased scattering
mogenization. This type of structural defects sets the glassy system far losses in this spectral region, this kind of structural disordering being
away from thermodynamic equilibrium, because in addition to increased suppressed in As-rich g-As42S58, which is quite close to the boundary
atomic coordination for As atom in optimally-constrained QT environ- of glass-forming region.
ment (nc = 3.00), the appearance of homopolar As\As bonds introduces Medium-range structural disordering in g-As\S is probed with the
extra constraints in a glassy matrix making its rigid and stress. Within FSDP-related XRD. The most essential influence expressed in the in-
constraint counting algorithm developed at the basis of the Phillips– creased effective periodicity R and correlation length L corresponding
Thorpe mean-field rigidity percolation theory [64,65], it can be easily to the FSDP is characterized for stoichiometric g-As40S60 prepared
estimated that the _As\[As(S2/2)]_S structural fragment having in the LT-regime. This specimen obviously demonstrates composition-
an adjacent As\As bond is over-constrained with n c = 3.20, al misbalance tending towards As-rich glass-forming network ap-
while _As\[As(S3/2)]\[As(S2/2)]_S unit with a more distinct As\As proaching g-As42S58.
bond has nc = 3.13. Thus, the coordination disordering results in This work gives first experimental evidence for microstructure
over-constrained glassy network in respect to chemical bond dis- changes corresponding to such inhomogeneities in S distribution
proportionality given by the following reaction below: using micro-Raman scattering. Owing to slow cooling from relatively
low temperatures and insufficient melt homogenization, the double-
3  S2=2 As–S–AsS2=2 ↔2  S3=2 As ¼ S þ 2  S2=2 As–AsS2=2 : ð7Þ
bond-based S3/2As_S units can be formed in g-As\S as a specific kind
of coordination disordering associated with non-uniform S distribution
The left-hand part of this equation represents the network of in a glassy matrix. Possible topological configurations of these structural
corner-shared AsS3/2 pyramids, while the right-hand side corresponds anomalies are examined with ab-initio quantum chemical modeling for
to QT S3/2As_S units and homopolar As\As bonds within As2S4/2 con- clusters, evolving both double-bond QT unit and neighboring S-
figuration. Under conventional synthesis conditions, the equilibrium of compensating homopolar As\As bond. Effect of coordination dis-
this reaction (6) is shifted in the left side, which determines the prefer- ordering, which is most pronounced in near-stoichiometric g-As40S60,
ence of heteropolar covalent bonding in g-As\S. However, under insuffi- does not result in optimally-constrained glassy network in a sense of
cient melt homogenization, which occurs at relatively low temperatures global connectivity.
Tq (450–500 °C), the bond disproportionality in the quenched glass at-
tains an opposite trend, producing fourfold-coordinated As atoms and References
“wrong” homopolar As\As bonds.
Supposedly, the coordination disordering could result in optimally- [1] A. Feltz, Amorphous Inorganic Materials and Glasses, VCH Publ, New York, 1993.
[2] Z.U. Borisova, Glassy Semiconductors, Plenum Press, New York-London, 1981.
constrained glassy network in a sense of global connectivity, provided
[3] A. Zakery, S.R. Elliott, Optical Non-linearities in Chalcogenide Glasses and Their Ap-
As\As bonds were overcompensated by S additions to form the plications, Springer-Verlag, Berlin-Heidelberg, 2007.
normally-coordinated covalent network of corner-sharing AsS3/2 pyra- [4] J.S. Sanghera, I.D. Aggarwal, Active and passive chalcogenide glass optical fibers for
mids. However, as it follows from Fig. 4, this effect determined in the in- IR applications: a review, J. Non-Cryst. Solids 256–257 (1999) 6–16.
[5] B. Bureau, X. Zhang, F. Smektala, J.-L. Adam, J. Troles, H. Ma, C. Boussard-Pledel, J.
creased stretching vibrations of QT units (530–540 cm−1) decays with S Lucas, P. Lucas, D. Le Coq, M.R. Riley, J.H. Simmons, Recent advances in chalcogenide
addition (in g-As38S62-LT) as compared with stoichiometric g-As2S3, glasses, J. Non-Cryst. Solids 345–346 (2004) 276–283.
O. Shpotyuk et al. / Journal of Non-Crystalline Solids 402 (2014) 236–243 243

[6] X. Zhang, B. Bureau, C. Boussard-Pledel, J. Lucas, P. Lucas, Glasses for seeing beyond [38] R. Holomb, M. Veres, V. Mitsa, Ring-, branchy-, and cage-like AsnSm nanoclusters in
visible, Chem. Eur. J. 14 (2008) 432–442. the structure of amorphous semiconductors: ab initio and Raman study, J.
[7] B.J. Eggleton, Chalcogenide photonics: fabrication, devices and applications, Opt. Ex- Optoelectron. Adv. Mater. 11 (2009) 917–923.
press 18 (2010) 26632–26634. [39] C.Y. Yang, D.E. Sayers, M.A. Paesler, X-ray-absorption spectroscopy studies of glassy
[8] K. Tanaka, Chemical and medium-range orders in As2S3 glass, Phys. Rev. B36 (1987) As2S3. The role of rapid quenching, Phys. Rev. B36 (1987) 8122–8128.
9746–9752. [40] S.L. Kuznetsov, M.D. Mikhailov, I.M. Pecheritsyn, E.Yu. Turkina, Structural chemical
[9] K. Tanaka, Interrelations between optical absorption edges and structural order in processes at the synthesis of chalcogenide glasses, J. Non-Cryst. Solids 213–214
glassy As2S3, Solid State Commun. 56 (1985) 899–903. (1997) 68–71.
[10] R. Golovchak, A. Kozdras, O. Shpotyuk, Optical signature of structural relaxation in [41] S. Mamedov, A. Kisliuk, D. Quitmann, Effect of preparation conditions on the low
glassy As2S3, J. Non-Cryst. Solids 356 (2010) 1149–1152. frequency Raman spectrum of glassy As2S3, J. Mater. Sci. 33 (1998) 41–43.
[11] M. Hansen, K. Anderko, Constitution of Binary Alloys, McGraw-Hill, New York, 1958. [42] K.S. Andrikopoulos, A.G. Kalampounias, O. Falagara, S.N. Yannopoulos, The glassy
[12] G.Z. Vinogradova, Glass Forming and Phase Equilibriums in Chalcogenide Systems, and supercooled state of elemental sulfur: vibrational modes, structure metastabil-
Nauka, Moscow, 1984. (in Russ.). ity, and polymer content, J. Chem. Phys. 139 (2013) 124501-1–124501-11.
[13] P. Bohac, A. Della Casa, A. Gaumann, Kristallziichtung von Arsentrisulfid, Krist. Tech. [43] R. Steudel, H.-J. Mausle, Molekulare Zusammensetzung von flussigem Schwefel. Teil
9 (1974) 237–241. 2: Qualitative Analyse und Isolierung von S7, S12, α-S18 und S20, Z. Anorg. Chem. 478
[14] R. Steudel, B. Eckert, Topics in Current Chemistry, Springer-Verlag, Berlin-Heidelberg, (1981) 156–176.
2003. [44] C. Liu, G. Tang, L. Luo, W. Chen, Phase separation inducing controlled crystallization
[15] L.A. Clark, The Fe–As–S system — phase relations and applications, Econ. Geol. 55 of GeSe2–Ga2Se3–CsI glasses for fabricating infrared transmitting glass-ceramics, J.
(1969) 1345–1381. Am. Ceram. Soc. 92 (2009) 245–248.
[16] F.A. Shunk, Constitution of Binary Alloys, Second Supplement, McGraw-Hill, New [45] C.F. Bohren, D.R. Huffman, Absorption and Scattering of Light by Small Particles, John
York, 1969. Wiley & Sons, New York, 1983.
[17] J.S. McCloy, B.J. Riley, S.K. Sundaram, H.A. Qiao, J.V. Crum, B.R. Johnson, Structure- [46] M. Yamaguchi, The relationship between optical gap and chemical composition in
optical property correlations of arsenic sulfide glasses in visible, infrared, and sub- chalcogenide glasses, Phil. Mag. B51 (1985) 651–663.
millimeter regions, J. Non-Cryst. Solids 356 (2010) 1288–1293. [47] S. Chakravarty, D.G. Georgiev, P. Boolchand, M. Micoulaut, Ageing, fragility and the
[18] R. Golovchak, O. Shpotyuk, J.S. McCloy, B.J. Roley, C.F. Windisch, S.K. Sundaram, A. reversibility window in bulk alloy glasses, J. Phys. Condens. Matter 17 (2005) L1–L7.
Kovalskiy, H. Jain, Structural model of homogeneous As–S glasses derived from [48] R. Steudel, K. Bergemann, J. Buschmann, P. Luger, Application of dicyanohexasulfane
Raman spectroscopy and high-resolution XPS, Phil. Mag. 90 (2010) 4489–4501. for the synthesis of cyclo-nonasulfur: crystal and molecular structures of S6(CN)2
[19] S.K. Sundaram, J.S. McCloy, B.J. Riley, M.K. Murphy, H.A. Qiao, C.F. Windisch, E.D. and of α-S19, Inorg. Chem. 35 (1996) 2184–2188.
Walter, J.V. Crum, R. Golovchak, O. Shpotyuk, Gamma radiation effects on physical, [49] B. Meyer, Elemental sulfur, Chem. Rev. 76 (1976) 367–389.
optical, and structural propertyies of binary As–S glasses, J. Am. Ceram. Soc. 95 [50] P. Chen, C. Holbrook, P. Boolchand, D.G. Georgiev, K.A. Jackson, M. Micoulaut, Inter-
(2011) 1048–1055. mediate phase, network demixing, boson and floppy modes, and compositional
[20] S.R. Elliott, Extended-range order, interstitial voids and the first sharp diffraction trends in glass transition temperatures of binary AsxS1 − x system, Phys. Rev. B78
peak of network glasses, J. Non-Cryst. Solids 182 (1995) 40–48. (2008) 224208-1–224208-15.
[21] S.G. Bishop, N.J. Shevchik, Small angle X-ray scattering evidence for the absence of [51] D.G. Georgiev, P. Boolchand, K.A. Jackson, Intrinsic nanoscale phase separation of
voids in chalcogenide glasses, Solid State Commun. 15 (1974) 629–633. bulk As2S3 glass, Phil. Mag. 83 (2003) 2941–2953.
[22] L.E. Busse, S.R. Nagel, Temperature dependence of the structure factor of As2Se3 [52] A. Kovalskiy, J. Cech, M. Vlcek, C.M. Waits, M. Dubey, W.R. Haffner, H. Jain, Chalco-
glass up to the glass transition, Phys. Rev. Lett. 47 (25) (1981) 1848–1851. genide glass e-beam and photoresists for ultrathin grayscale pattering, J. Micro/
[23] L.E. Busse, Temperature dependence of the structures of As2Se3 and AsxS1 − x glasses Nanolithogr. MEMS MOEMS 8 (2009) 043012-1–043012-11.
near the glass transition, Phys. Rev. B 29 (6) (1984) 3639–3651. [53] R. Holomb, V. Mitsa, P. Johansson, N. Mateleshko, A. Matic, M. Veresh, Energy-
[24] K. Tanaka, Pressure dependence of the first sharp diffraction peak in chalcogenide dependence of light-induced changes in g-As45S55 during recording the micro-
and oxide glasses, Philos. Mag. Lett. 57 (1988) 183–187. Raman spectra, Chalcogenide Lett. 2 (2005) 63–69.
[25] L. Cervinka, Medium-range ordering in non-crystalline solids, J. Non-Cryst. Solids 90 [54] F. Billes, V. Mitsa, I. Fejes, N. Mateleshko, J. Fejsa, Calculation of the vibrational spec-
(1987) 371–382. tra of arsenic sulfide clusters, J. Mol. Struct. 513 (1999) 109–115.
[26] S.R. Elliott, Origin of the first sharp diffraction peak in the structure factor of covalent [55] M. Becucci, R. Bini, E. Castellucci, B. Eckert, H.J. Jodl, Mode assignment of sulfur α-S8
glasses, Phys. Rev. Lett. 67 (1991) 711–714. by polarized Raman and FTIR studies at low temperatures, J. Phys. Chem. B101
[27] M. Shpotyuk, O. Shpotyuk, R. Golovchak, P. Demchenko, FSDP-related correlations in (1997) 2132–2137.
γ-irradiated chalcogenide semiconductor glasses: the case of glassy arsenic trisul- [56] J. Kolar, L. Strizik, T. Kohutek, T. Wagner, G.A. Voyiatzis, A. Chrissanthopoulos, S.N.
fide g-As2S3 revised, J. Phys. Chem. Solids 74 (2013) 41721–41725. Yannopoulos, Influence of thermal history on the photostructural changes in glassy
[28] M. Shpotyuk, O. Shpotyuk, R. Serkiz, P. Demchenko, S. Kozhyukhin, Surface oxida- As15S85 studied by Raman scattering and ab initio calculations, J. Appl. Phys. 114
tion in glassy arsenic trisulfide induced by high-energy γ-irradiation, Radiat. Phys. (2013) 203502-1–203502-7.
Chem. 97 (2014) 341–345. [57] V. Boyko, O. Shpotyuk, M. Hyla, Double-bond defect modelling in As–S glasses, Inst.
[29] T. Roisnel, J. Rodriguez-Carvajal, WinPLOTR: a Windows tool for powder diffraction Phys. Conf. Ser. 15 (2010) 012074-1–012074-4.
patterns analysis, Mater. Sci. Forum 118 (2001) 378–381. [58] P.H. Gaskell, The structure of simple glasses: randomness or pattern the debate goes
[30] Yu.I. Sozin, Diffractometry of coordination spheres, Crystallogr. Rep. 39 (1994) 6–13. on, Glas. Phys. Chem. 24 (1998) 180–187.
[31] O. Ehrenfest, On interference phenomena to be expected when Röntgen rays pass [59] A.C. Wright, Longer range order in single component network glasses? Phys. Chem.
through a diatomic gas, Proc. Royal Acad. Amsterdam XVII (1915) 1184–1190. Glasses Eur. J. Glass Sci. Technol. B49 (2008) 103–117.
[32] O. Shpotyuk, V. Boyko, M. Hyla, Cation-interlinking network cluster approach in [60] W.H. Zachariasen, The atomic arrangement in glass, J. Am. Chem. Soc. 54 (1932)
application to extended defects in covalent-bonded glassy semiconductors, Phys. 3841–3851.
Status Solidi C 6 (2009) 1882–1885. [61] E. Bychkov, C.J. Benmore, D.L. Price, Compositional changes of the sharp diffraction
[33] O. Shpotyuk, M. Hyla, V. Boyko, Structural-topological genesis of network-forming peak in binary selenide glasses, Phys. Rev. B 72 (2005) 172107-1–172107-4.
nanoclusters in chalcogenide semiconductor glasses, J. Optoelectron. Adv. Mater. [62] R. Golovchak, H. Jain, O. Shpotyuk, A. Kozdras, A. Saiter, J.-M. Saiter, Experimental
15 (2013) 1429–1437. verification of the reversibility window concept in binary As–Se glasses subjected
[34] W.J. Hehre, R.F. Stewart, J.A. Pople, Self-consistent molecular orbital methods I. Use to a long-term physical aging, Phys. Rev. B 78 (2008) 014202-1–014202-6.
of Gaussian expansions of Slater type atomic orbitals, J. Chem. Phys. 51 (1969) [63] R. Davies, Chalcogen-phosphorus (and heavier congeners) chemistry, in: F.A.
2657–2665. Devillanova (Ed.), Handbook of Chalcogen Chemistry. New Perspectives in Sulfur,
[35] A.D. McLean, G.S. Chandler, Contracted Gaussian basis sets for molecular calcula- Selenium and Tellurium, RSC Publ, Cambridge, 2007, pp. 286–336.
tions. I. Second row atoms, Z = 11–18, J. Chem. Phys. 72 (1980) 5639–5648. [64] J.C. Phillips, Topology of covalent non-crystalline solids. I: short-range order in chal-
[36] K. Jackson, A. Briley, S. Grossman, D. Porezag, M. Pederson, Raman-active modes of α- cogenide alloys, J. Non-Cryst. Solids 34 (1979) 153–181.
GeSe2 and α-GeS2: a first-principles study, Phys. Rev. B60 (1999) R14985–R14989. [65] M.F. Thorpe, Continues deformations in random networks, J. Non-Cryst. Solids 57
[37] K. Jackson, Electric fields in electronic structure calculations: electric polarizabilities (1983) 355–370.
and IR and Raman spectra from first principles, Phys. Stat. Sol. B217 (2000)
293–310.

You might also like