Next Article in Journal
Effect of Apis mellifera syriaca Bee Venom on Glioblastoma Cancer: In Vitro and In Vivo Studies
Previous Article in Journal
Modification of Fungicide Treatment Needs and Antioxidant Content as a Result of Real-Time Ozonation of Raspberry Plants
Previous Article in Special Issue
Why Do Ionic Surfactants Significantly Alter the Chemiluminogenic Properties of Acridinium Salt?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nanostructured Molecular–Network Arsenoselenides from the Border of a Glass-Forming Region: A Disproportionality Analysis Using Complementary Characterization Probes

by
Oleh Shpotyuk
1,2,*,
Malgorzata Hyla
1,
Adam Ingram
3,
Yaroslav Shpotyuk
4,5,
Vitaliy Boyko
2,
Pavlo Demchenko
6,
Renata Wojnarowska-Nowak
7,
Zdenka Lukáčová Bujňáková
8 and
Peter Baláž
8
1
Institute of Physics, Jan Dlugosz University in Częstochowa, 13/15, al. Armii Krajowej, 42-200 Częstochowa, Poland
2
O.G. Vlokh Institute of Physical Optics, Ivan Franko National University of Lviv, 23, Dragomanov Str., 79005 Lviv, Ukraine
3
Faculty of Physics, Opole University of Technology, 75, Ozimska Str., 45370 Opole, Poland
4
Department of Sensor and Semiconductor Electronics, Ivan Franko National University of Lviv, 107, Tarnavskoho Str., 79017 Lviv, Ukraine
5
Institute of Physics, University of Rzeszow, 1, Pigonia Str., 35-959 Rzeszow, Poland
6
Department of Inorganic Chemistry, Ivan Franko National University of Lviv, 6, Kyryla i Mefodiya Str., 79000 Lviv, Ukraine
7
Center for Microelectronics and Nanotechnology, Institute of Materials Engineering, University of Rzeszow, 1, Pigonia Str., 35-959 Rzeszow, Poland
8
Institute of Geotechnics of Slovak Academy of Sciences, 45, Watsonova Str., 04001 Košice, Slovakia
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(16), 3948; https://doi.org/10.3390/molecules29163948
Submission received: 25 July 2024 / Revised: 13 August 2024 / Accepted: 14 August 2024 / Published: 21 August 2024
(This article belongs to the Special Issue Exclusive Feature Papers in Physical Chemistry, 2nd Edition)

Abstract

:
Binary AsxSe100−x alloys from the border of a glass-forming region (65 < x < 70) subjected to nanomilling in dry and dry–wet modes are characterized by the XRPD, micro-Raman scattering (micro-RS) and revised positron annihilation lifetime (PAL) methods complemented by a disproportionality analysis using the quantum–chemical cluster modeling approach. These alloys are examined with respect to tetra-arsenic biselenide As4Se2 stoichiometry, realized in glassy g-As65Se35, glassy–crystalline g/c-As67Se33 and glassy–crystalline g/c-As70Se30. From the XRPD results, the number of rhombohedral As and cubic arsenolite As2O3 phases in As-Se alloys increases after nanomilling, especially in the wet mode realized in a PVP water solution. Nanomilling-driven amorphization and reamorphization transformations in these alloys are identified by an analysis of diffuse peak halos in their XRPD patterning, showing the interplay between the levels of a medium-range structure (disruption of the intermediate-range ordering at the cost of an extended-range one). From the micro-RS spectroscopy results, these alloys are stabilized by molecular thioarsenides As4Sen (n = 3, 4), regardless of their phase composition, remnants of thioarsenide molecules destructed under nanomilling being reincorporated into a glass network undergoing a polyamorphic transition. From the PAL spectroscopy results, volumetric changes in the wet-milled alloys with respect to the dry-milled ones are identified as resulting from a direct conversion of the bound positron–electron (Ps, positronium) states in the positron traps. Ps-hosting holes in the PVP medium appear instead of positron traps, with ~0.36–0.38 ns lifetimes ascribed to multivacancies in the As-Se matrix. The superposition of PAL spectrum peaks and tails for pelletized PVP, unmilled, dry-milled, and dry–wet-milled As-Se samples shows a spectacular smoothly decaying trend. The microstructure scenarios of the spontaneous (under quenching) and activated (under nanomilling) decomposition of principal network clusters in As4Se2-bearing arsenoselenides are recognized. Over-constrained As6·(2/3) ring-like network clusters acting as pre-cursors of the rhombohedral As phase are the main products of this decomposition. Two spontaneous processes for creating thioarsenides with crystalline counterparts explain the location of the glass-forming border in an As-Se system near the As4Se2 composition, while an activated decomposition process for creating layered As2Se3 structures is responsible for the nanomilling-driven molecular-to-network transition.

1. Introduction

The binary arsenoselenides of an AsxSe100−x cut-section comprise an important family of substances with saturated covalent bonding [1,2,3,4], where glass-like specimens can easily be obtained via high-entropy melting by quenching (melt-quenching, MQ) within an unprecedentedly wide compositional domain 0 ≤ x < ~70 corresponding to coordination numbers (CN) 2.0 < CN < ~2.70 [1,2]. Because of their vitreous state, these materials represent promising candidates for multifunctional applications in electronics, optoelectronics, IR optics, chemical and biochemical sensing, etc. [4,5,6]. In the early 2010s, the achieved progress in such implementation spheres was so impressive that a new branch of IR photonics, nominated Chalcogenide Photonics, appeared [7,8].
Nowadays [9,10,11], it is obvious that the most important functionality of arsenoselenides is governed by the specificity of their vitreous state, which can be modified either intrinsically, owing to the optimized technology of MQ, or extrinsically, owing to the post-technological influence on MQ-derived alloys. As for the latter, essential progress has been achieved in the most recent decade on so-called nanostructurization routines like high-energy mechanical milling (MM), also termed nanomilling [12], applied to transfer these substances from the macro- to nanoscopic scales, exploring the unique possibilities of their newly attained nanofunctionality [13,14,15,16,17]. From this point, the As-bearing arsenoselenides beyond stoichiometric arsenic triselenide As2Se3 (x > 40), which possess glass-like structures due to the thioarsenide As4Sen molecules stabilized in the network of remaining As-Se bonds, are especially attractive [1,2,3]. In this research, the thioarsenide nomenclature As4Xn recommended for As-rich over-stoichiometric sulphides (X = S) under an analysis of the molecular-packing conformations by Bonazzi and Bindi [18] and the electron density by Gibbs et al. [19] is used for AsxSe100−x alloys. A milling-driven escape towards an optimized nanostructured state occurs in these substances via transitions between the amorphous states, referred to as polyamorphic or reamorphization transitions, similar to the characteristics of chemically stoichiometric arsenic triselenide As2Se3 (corresponding to x = 40 in AsxSe100−x, CN = 2.40) and under-stoichiometric Se-rich glassy alloys (0 < x < 40; 2.0 < CN < 2.40) [20,21]. In contrast, in compositional domains where some molecular species can be stabilized in a glass-forming network, transitions between different crystalline and amorphous states (referred to as polymorphic and polyamorphic phase transitions) can be activated by nanomilling in a dry mode, like in Se-rich AsxSe100−x glassy alloys (x < 40), approaching ‘pure’ Se (0 ≤ x < ~10–12) [22] or As-rich alloys (x > 40) of arsenic monoselenide, g-AsSe (also referred to as tetra-arsenic tetraselenide, g-As4Se4) [23]. Although such transitions have a higher content of side-oxidized derivatives, they can be activated in alloys subjected to nanomilling in a wet and/or combined dry–wet mode, typically realized in an environment with some surfactants, such as the PVP (polyvinylpyrrolidone) water solution [24,25,26].
Specifically, it is expected that both the polymorphic and polyamorphic phase transitions can be activated in the MQ-derived AsxSe100−x alloys at the border of the glass-forming region ~(65 < x < 70) enriched in glassy and glassy–crystalline species undergoing nanostructurization by means of MM, with this process being dependent on the MM mode (dry or wet). There are no crystalline counterparts to molecular thioarsenides at the border of the glass-forming region in an As-Se system such as tetra-arsenic biselenide As4Se2 (corresponding to CN = 2.67, which is just at the edge of the glass-forming ability in this system [1,2]), while some thioarsenide molecules in a nearby region (such as tetra-arsenic triselenide As4Se3 with CN = 2.67, or tetra-arsenic monoselenide As4Se with CN = 2.80) are still expected under grinding in a high-energy ball mill, serving as a source of possible network-forming derivatives contributing to the final state of these nanostructured substances.
The scope of this work is comprehensive research on the AsxSe100−x alloys from the border of the glass-forming region (2.65 < CN < 2.70) subjected to nanomilling in dry- and combined dry–wet modes, using multi-experimental characterization probes like X-ray powder diffraction (XRPD) analysis in terms of a modified microcrystalline model [20,21,22,23], micro-Raman scattering (micro-RS) spectroscopy [27] and revised positron annihilation lifetime (PAL) analysis within the Positronics approach [28,29], complemented with molecular–network disproportionality analysis in the As-Se system employing ab initio quantum–chemical cluster-modeling algorithm CINCA like in [22,23]. In the current research, the compositional domain of interest in the As-Se system, which will be shifted towards the border of the glass-forming region, exemplified by arsenoselenide alloys deviated around tetra-arsenic biselenide As4Se2 thioarsenide stoichiometry, these being under-stoichiometric glassy g-As65Se35 (CN = 2.65), stoichiometric glassy-crystalline g/c-As67Se33 (viz. g/c-As4Se2, CN = 2.67) and over-stochiometric glassy-crystalline g/c-As70Se30 (CN = 2.70).

2. Results and Discussion

2.1. Atomic-Specific Microstructure of the Examined Arsenoselenides

The normalized XRPD profiles of unmilled and dry-milled g/c-As67Se33 alloy (CN = 2.67) strictly corresponding to tetra-arsenic biselenide (As4Se2) thioarsenide stoichiometry along with the Bragg diffraction reflexes from possible As-Se crystalline counterparts [30,31,32,33,34,35,36,37] are reproduced in Figure 1. These XRPD patterns demonstrate a so-called three-peak amorphous structure of this substance composed of three principal diffuse peak halos responsible for the FSDP (the first sharp diffraction peak) at ~15–25°2θ, SSDP (the second sharp diffraction peak) at ~28–33°2θ and TDP (the third diffraction peak) at ~50–60o2θ, superimposed with sharp-broadened Bragg diffraction reflexes from nanocrystalline inclusions of elemental As [30,31] and its oxide As2O3 [38]. In part, analysis of the XRPD patterns reproduced in Figure 1 and Figure 2 for g/c-As67Se33 samples of different pre-history confirms that the strongest diffraction lines arise from the planes (110), (0–11) and (211) in rhombohedral As [30,31], and (222), (111), (331), (400) and (440) in cubic arsenolite As2O3 [38].
In unmilled g/c-As67Se33 (Figure 2a), the elemental As is dominated over its oxide (As2O3) due to the sharp ‘nanocrystalline’ reflexes of this phase superimposed on relatively weak ‘amorphous’ halos, the FSDP-related diffuse peak halo being equilibrated by intensity with that responsible for the SSDP at ~28–33°2θ, which is characteristic of vitreous selenides [39,40].
In dry-milled g/c-As67Se33 treated in a protective argon atmosphere (Figure 2b), the diffuse peak halos prevail over broadened ‘nanocrystalline’ reflexes from elemental As and its oxide, confirming that strong amorphous-I-to-amorphous-II (reamorphization) transition in this sample is probably accompanied by crystalline-to-amorphous (amorphization) transition, as this occurs in g-AsSe subjected to high-energy MM in a dry mode [23]. All three principal diffuse peak halos in dry-milled g/c-As67Se33 fit in a decaying sequence on their intensities like in Sulphur S-bearing glasses [39], meaning that stress generated under grinding does not completely relax over the glass network.
Then, in dry-milled g/c-As67Se33 additionally treated in a wet mode (that is, in dry–wet-milled samples, see Figure 2c), ‘nanocrystalline’ lines of As2O3 phase abnormally grow in intensity and become broader in width over those of elemental As, and the arrangement of principal peak halos corresponding to the amorphous phase attains a characteristic irregularity, which dominated the SSDP-related peak halo at ~28–33°2θ as a signature of thermodynamically stabilized glassy alloys from the glass-forming region [20,21,22,23]. Therefore, it seems quite reasonable that MM in the PVP water solution promotes further extraction and oxidation of the As phase. In the g-As65Se35 sample from the edge of the glass-forming region, which was in the vitreous state before and after dry MM [26], nanomilling in a wet mode results in an arsenolite As2O3 phase without a notable admixture of the unoxidized elemental As phase (Figure 3). In line with our expectation, this oxidation process is enhanced under the transition to glassy-crystalline alloys from the border of the glass-forming region such as g/c-As70Se30 (see Figure 3).
The MM-driven amorphization and reamorphization processes in the examined arsenoselenides can be adequately identified due to the analysis of the arrangement of diffuse peak halos in their XRPD patterns reproduced in Figure 1, Figure 2 and Figure 3, the derived FSDP parameters related to intermediate-range ordering in these alloys being summarized in Table 1.
From a detailed inspection of Figure 1, the position of the FSDP-related peak halo in unmilled g/c-As67Se33 at 15.479°2θ (equivalent to characteristic distance in the Bragg diffraction R~5.71 Å, or average inter-atomic distance in the Ehrenfest diffraction ds = 7.02 Å) agrees well with the most intensive diffraction line (I = 100%) in orthorhombic As4Se3 which arise from the (111) plane at ~16.90°2θ (equivalent to inter-planar distance d = 5.243 Å [32]). One of the strongest (I = 91.3%) lines in other molecular thioarsenide (monoclinic As4Se4 [33,34]) arises from the (120) plane at ~16.07°2θ (d = 5.512 Å), which is also close to the FSDP position. The other line (I = 91.2%) ascribed to the (020) plane in monoclinic As2Se3 [35,36] is revealed at higher diffraction angles ~17.9°2θ (d = 4.950 Å). Assuming equal contributions from these inter-planar correlations, the FSDP-related characteristic distance R can be found to approach ~5.64 Å, instead of ~5.71 Å from solely the Bragg diffraction positioning of this diffuse peak halo (see Table 1). So, another input to the FSDP is expected from the Ehrenfest diffraction contributing through inter-atomic and/or inter-molecular correlations, which belong to molecular thioarsenide structures and their network-forming derivatives, with average inter-atomic/molecular distances ds beyond ~7 Å [23].
In As-rich g-AsxSe100−x (40 < x < 65) [21], remnants of thioarsenide cage-like As4Sen molecules (such as As4Se4, As4Se3, As4) are known to be destroyed under MM and re-incorporated into a glass network resulting in an increased FSDP position Q1 and width ΔQ1. This amorphization trend corresponding to the crystalline-to-amorphous transition causes a fragmentation impact on the correlation length of the FSDP-responsible entities (L), facilitating changes when distant inter-atomic correlations between some crystalline planes contributing to the FSDP disappear at a cost of others responsible for the SSDP. The disruption of intermediate-range ordering due to weakening of the FSDP-responsible entities accompanied by an enhancement of extended-range ordering due to fragmentation of the SSDP-responsible entities occurs as interplay between different hierarchical levels of the medium-range structure in As-Se alloys from the glass-forming region subjected to MM [21].
Similar peculiarities are found in the examined glassy and glassy-crystalline As-Se alloys from the border of the glass-forming region. Thus, in dry-milled alloys, the FSDP position Q1 slightly shifts in the high-angular side, while the FSDP width ΔQ1 obeys a drastic increase, resulting in fragmentation of the correlation length L (see Table 1). These changes prevail in glassy-crystalline substances such as g/c-As70Se30, confirming that crystalline remainders serve as a possible source for the MM-driven amorphization (crystalline-to-amorphous) transition. As shown in Figure 2, nanomilling in combined dry–wet mode (when the last processing stage is performed in the PVP water solution) increases the breakdown in intermediate-range ordering. Nanomilling in wet mode does not change the principal appearance of diffuse peak halos in the XRPD patterns but leads to drastic broadening in their widths, in favor of enhanced reamorphization in preliminary dry-milled samples.
Thus, remnants of ‘crystalline’ structures in the examined alloys responsible for inter-molecular correlations with inter-centroid distances between thioarsenide cage-like molecules above ~7 Å (which contribute to the XRPD patterning by the Ehrenfest diffraction) are destroyed under MM, while the inter-planar correlations contribute further to these patterns through Bragg diffraction. Because of molecular-to-network reamorphization transformations in the examined As-Se alloys, the broadened and depressed diffuse peak halos become shifted towards higher scattering vectors. Since remnants of thioarsenide As4Sen molecular entities destructed under grinding interact with oxygen, especially under MM in the PVP water solution, the arsenolite (As2O3) phase is formed and stabilized in these alloys subjected to nanomilling in combined dry–wet mode.
To shed more light on the microstructure transformations in the examined arsenoselenide alloys, the micro-RS spectra were collected for these samples before and after the respective stages of their processing. For g/c-As67Se33 samples taken as unmilled bulk pieces and pelletized coarse-grained, dry-milled and dry–wet-milled samples, these micro-RS spectra normalized with respect to the amplitude of the dominant band are, respectively, reproduced in Figure 4a–d.
Firstly, we can confirm that palletization under the above conditions did not affect the micro-RS spectra of these samples. Indeed, this conclusion becomes obvious from a visual comparison of these spectra in Figure 4a,b collected for bulk pieces and pellets of MQ-derived g/c-As67Se33. Further, specifically for the Raman-active bands contributing to these spectra, a few low-frequency (at ~130, 143, 154, 168 and 190 cm−1) and high-frequency (at ~190, 204, 218, 236, 253, 278 cm−1) bands are resolved in the normalized micro-RS spectrum of unmilled g/c-As67Se33 (see Figure 4a). The high-frequency bands are ascribed to overlapped (mainly strong) bond-stretching modes of AsSe3 pyramids at 220–230 cm−1 [3,41,42] and thioarsenide molecules, such as As4Se4 (peaks at ~225 and 240 cm−1 supplemented by shoulder near 255 cm−1 [3,42]), As4Se3 (peaks at ~195, 205, 225, 240 and 255 cm−1 supplemented by shoulder near 280 cm−1 [42]) and As4 (strong peak at ~200–210 cm−1 [3,43,44]). The low-frequency bands (preferentially weak and medium) can be ascribed to the bond-bending modes of thioarsenide molecules, in part, As4Se4 (at ~145, 160 and 170 cm−1 [42,45,46]) and As4Se3 (at ~110, 135, 150, 160 and 170 cm−1 [42,45]). Thus, with respect to micro-RS spectroscopy, the appearance of the As phase is expected in MQ-derived g/c-As67Se33 samples due to the RS-active band at ~204 cm−1 ascribed to As4 thioarsenide molecules, while both As4Se4 and As4Se3 molecules are revealed via the strong band at ~236 cm−1 supplemented by a weak shoulder at ~278 cm−1 (see Figure 4a,b). The spectral region between these bands (220–230 cm−1) corresponds to the vibration modes of AsSe3 pyramids incorporated in a network via = As-Se-As = bridges. As a result, in g/c-As67Se33, the RS spectrum composed of overlapped high-frequency modes of thioarsenide molecules and network-forming fragments obeys a double-peak-shape characteristic with strong maxima at ~200 and ~240 cm−1 supplemented by a relatively suppressed hump at ~220–230 cm−1 and weak shoulder near ~280 cm−1 (for a comparison, see [3,27,42,45,46]).
These features revealed in micro-RS spectra of unmilled g/c-As67Se33 (shown in Figure 4a,b) are essentially broadened in this sample subjected to nanomilling in dry mode (see Figure 4c) or combined dry–wet mode (see Figure 4d). Such behavior speaks in favor of a nanomilling-driven molecular-to-network amorphization transition, which occurs due to the destruction of thioarsenide molecules followed by the incorporation of their remnants into the glass network undergoing a polyamorphic transition (amorphous-I-to-amorphous-II [47,48] or, alternatively, reamorphization transition [23]). These alloys, affected by defects generated under MM, become notably stressed in the nanomilled state, enhancing the calorimetric heat transfer responses, as compared with unmilled bulk samples [27].
Noteworthily, the micro-RS spectroscopic identification of the rhombohedral (grey) As phase, which is expected due to two-fold degenerate Eg and A1g modes near ~202 cm−1 and 254 cm−1 [49], is impossible in these alloys because of overlapping with the vibration bands of the arsenoselenide network. The same is true of nanocrystalline inclusions of the cubic arsenolite (As2O3) phase, which is expected in these alloys due to RS-active modes near ~190 cm−1 and ~270 cm−1 [44,49], which are also overlapped with As-Se network modes in this spectral range. Therefore, the absence of the above crystalline inclusions in the examined alloys does not change the principal appearance of their micro-RS spectra, as demonstrated in Figure 5 for purely glassy-like specimens (g-As65Se35) studied before and after dry MM.
Thus, atomic-specific microstructure transformations in the examined As-Se alloys driven by nanomilling are dominated by the polyamorphic (molecular-to-network) transitions of the existing MQ-derived amorphous phase (forming reamorphization channel), supplemented (especially under MM in a dry mode) by weak polymorphic (crystalline-to-amorphous) transitions of crystalline inclusions stabilized in these alloys under MQ (forming amorphization channel). Undoubtedly, because of the principal difference in the atomic packing of amorphous and crystalline phases in arsenoselenide alloys undergoing nanostructurization by MM, such transformations should be accompanied by the notable changes in atomic-deficient (free-volume) microstructure of these alloys.

2.2. Atomic-Deficient Microstructure of the Examined Arsenoselenides

The raw PAL spectra reconstructed from unconstrained three-term fitting for g/c-As70Se30 alloys in the unmilled (bulky) state and after the respective stages of nanomilling are depicted in Figure 6, and they are compared with the spectrum of ‘pure’ PVP pelletized under the same conditions [50]. The applicability of this fitting procedure is proved by a scatter of variance (representing minimal statistically weighted least-square deviation between experimental and theoretical curve built of three exponentials) tightly grouped along the time axis in the bottom insets of Figure 6. Similar PAL spectra (not reproduced in Figure 6) were collected for other pelletized arsenoselenides. The best-fit PAL spectra parameters (component lifetimes τi and intensities Ii, I = 1, 2, 3) for arsenoselenide samples and PVP (pelletized under the same conditions [50]) are gathered in Table 2, and positron-trapping and Ps-decaying modes derived from these PAL spectra within two-state simple-trapping model (STM) [28,29] ignoring the Ps-decaying contribution are listed in Table 3.
Just from a visual inspection of Figure 6, it is clear that occupation of the ‘tail’ states related to bound positron–electron (positronium, Ps) states grows notably in bulk and dry-milled As-Se alloys under the transition to dry–wet-milled samples, approaching that in ‘pure’ PVP. This means that the transition from preferential annihilation via positron trapping to Ps decays. The PAL spectra of both unmilled and dry-milled arsenoselenides reveal a defect-specific lifetime of τ2~0.36–0.39 ns (Table 2), the value characteristic for multiatomic (bi-, tri-, quadruple) vacancies in amorphous As-Se [29,51,52,53,54,55]. The third component in the collected PAL spectra of these samples with τ3~2.1 ns and small fractional free volume below fv~0.3 can be ascribed to holes with a radius of R3~0.30 nm (Table 3). Because of the small input from the Ps-decay channel, the derived defect-free bulk lifetime τb approaches ~0.27 ns for the unmilled g-As65Se35 sample and somewhat lower, ~0.25 ns, for the g/c-As70Se30 sample. Noteworthily, the τb lifetime in crystalline arsenic triselenide is ~0.24 ns [51], which increased to ~0.28–0.29 ns in the amorphous matrix [29,52,53,54,55]. Therefore, positron trapping in unmilled and dry-milled arsenoselenides is expected to occur preferentially in the disordered As-Se network.
With the transition to samples pelletized from dry–wet-milled arsenoselenides (representing, in fact, PVP-capped nanocomposites [24,25,26]), the location of the preferential annihilation channel is changed. As follows from Table 2, the τ2 lifetime is notably increased in dry–wet-milled samples approaching ~0.43 ns (the value characteristic for positron trapping in large vacancies in As-Se matrix [51,52]), while the τ3 lifetime is decreased to ~1.9 ns (closer to Ps-related lifetime in PVP, τ3 = 1.867 ns [50]), resulting in an evidently increased average lifetime, τav.~0.38 ns. It means that additional free volume contributes to positron trapping sites in wet-milled samples while disappearing in Ps-hosting holes. Nevertheless, the more than three-times increased third component intensity (I3) speaks in favor of the fragmentation of Ps-decay sites, leading to fractional free volume fv~0.8 (Table 3). These changes are counterbalanced by decreased I2, preferring (under criterion [56]) an increased Ps-formation probability in the transition from dry-milled to dry–wet-milled arsenoselenide samples.
Therefore, the Ps decay in the PVP-capped nanocomposites occurs rather in ‘pure’ PVP environment than in the amorphous As-Se matrix. A comparative presentation of the respective PAL spectra for these arsenoselenide alloys and PVP shows a spectacular trend in their superposition, specifically the compositionally tuned behavior in the overlapped PAL spectra highlighted in Figure 7. Smooth monotonic changes are observed in both PAL spectra peaks and tails in these samples undergoing nanostructurization in the transition between states characteristic for unmilled (or dry-milled) and wet-milled samples. Similar behavior was also observed in other As-Se alloys from the glass-forming region [24,25,26]. In contrast, in the examined glassy and glassy-crystalline alloys from the border of the glass-forming region, an almost invariant tendency is observed in the PAL spectra peaks depressed on the right wing after wet milling due to the moderated Ps formation probability in PVP water solution and slightly changed average positron lifetime (τav). These changes in the PAL spectra tails for unmilled, dry-milled and dry–wet-milled samples are due to the increased density of Ps-hosting holes, defined merely by the I3 intensity. There is no evident empty gap between these spectra tails for dry-milled and dry–wet-milled g-As65Se35 samples (Figure 7a), as compared with glassy-crystalline g/c-As67Se33 or g/c-As70Se30 samples (see Figure 7b,c) caused by changes in the Ps-decaying channel under the transition from amorphous As-Se to the PVP-bearing environment. The richer variety of positron traps in glassy-crystalline alloys with τ2~0.43 ns (see Table 2) facilitates the formation of such an empty gap, as follows from the comparison between respective PAL spectra in Figure 7.
To shed more light on the examined As-Se alloys undergoing nanomilling-driven volumetric changes, the PAL spectra of dry-milled samples are considered with respect to these spectra in wet-milled samples employing revised positron lifetime analysis within the Positronics approach [28,29]. The trapping conversion parameters derived from such comparisons are presented in Table 4. Positive values of the first and second component intensities (In and Iint) in the reconstructed PAL spectrum of the heterogeneous medium undergoing volumetric nanostructurization, along with the well-defined component inputs, testify that Ps-hosting holes in these nanocomposites (g-As65Se35/PVP, g/c-As67Se33/PVP and g/c-As70Se30/PVP) are transformed in positron traps typical for dry-milled samples. The defect-specific lifetimes τint~0.36–0.38 ns (Table 4) show that positron traps are large interfacial free-volume voids close to the multiatomic vacancies in the amorphous As-Se matrix [51,52,53,54,55]. This finding agrees with (τint-τbNP) ~0.09–0.11 ns and τint/τbNP~1.4, indicative of large vacancies in chalcogenide compounds [57]. The overall conversion process, i.e., appearance of positron traps with τint~0.36–0.38 ns instead of Ps-hosting holes with τ3~2.1 ns, occurs in an environment with an effective bulk lifetime of τbNP~0.29 ns, which is above τb in crystalline arsenic selenide (~0.240 ns [51]) approaching the value characteristic for glassy As-Se (~0.28–0.29 ns) [52,53,54,55]. Therefore, we can reasonably speculate that a heterogeneous medium accommodates the products of direct Ps-to-positron conversion, i.e., unoccupied free-volume spaces where positron traps (typical for dry-milled As-Se alloys) appear instead of Ps-decaying holes (typical for PVP-stabilized nanocomposites) is the amorphous As-Se matrix, slightly enriched in some impurity products such as arsenic oxide.

2.3. Molecular–Network Disproportionality at the Border of the Glass-Forming Region in As-Se System via Quantum–Chemical Cluster Modelling

In substances like thioarsenide As4Sen-bearing compounds in the AsxSe100−x system, effects from the selective localization of positron-trapping and Ps-hosting sites are strongly enhanced due to the glassy-crystalline environment around intrinsic free-volumes elements (such as voids, atomic vacancies, vacancy-like clusters, pores, cracks, etc.) acting as potential setting places for trapped positrons and bound positron–electron (positronium, Ps) states [28,29]. Following the famous prediction of Blachnik and Wickel in 1984 on the thermal decomposition of A4B3 cage-like molecules into A4B4 ones and unidentified amorphous products [58], it has been proved recently [59] that nanomilling facilitates the decomposition of As4Se3 molecules in arsenoselenides within the glass-forming region in the vicinity of x = 57 (dimorphite-type composition), resulting in a reamorphized glass structure enriched in realgar-type As4Se4 molecules (prevailing near x = 50) and network remainders close to tetra-arsenic biselenide As4Se2 (dominated at x = 67). Since there are no molecular thioarsenides among products of devitrification in arsenoselenides from the border of the glass-forming region in the AsxSe100−x system (65 < x < 70), it seems reasonable to assume that crystallization processes are a result of the spontaneous (under MQ) or activated (under thermal or mechanical energy) decomposition of some network derivatives near As4Se2 composition (x = 67).
Cluster modelling of the molecular and network conformations was performed previously with the help of the ab initio quantum–chemical modelling algorithm CINCA [60,61] for AsxSe100−x alloys with crystalline counterparts, such as As2Se3 (x = 40) [62,63,64], As4Se4 (x = 50) [23] and As4Se3 (x = 57) [59]. On this basis, let us examine the forming energies (Ef) among As4Sen thioarsenide clusters (n = 6–0), covering the glass-forming border in the AsxSe100−x system (65 < x < 70), positioned around tetra-arsenic biselenide As4Se2 (x = 67, n = 2) and beyond this composition approaching tetra-arsenic As4 (x = 100, n = 0) and tetra-arsenic hexaselenide As4Se6 (x = 40, n = 6), the latter being equivalent to As2Se3.
In a group of tetra-arsenic hexaselenide (As4Se6) clusters [62,63,64], a strong network-forming trend clearly dominates, resulting in layered arsenoselenide 2D structures based on corner-shared AsSe3/2 pyramids with average cluster-forming energy, Ef ~ −72.619 kcal/mol [63], which is even better than the Ef energy of the single trigonal AsSe3/2 pyramid (−72.309 kcal/mol) [61]. The prevailing canonical chain-crossing model [1,2,3] in AsxSe100−x glasses at x < 40 is governed by decomposition on two-cation Se-rich clusters (under-constrained or floppy in view of number of constraints nc calculated after the Phillips–Thorpe algorithm [65,66,67] less than space dimensionality, 3.0) and As2Se3 clusters (optimally constrained, rigid but not stress, nc = 3.0) [61]. Disproportionality analysis shows that As2Se3-bearing network structures are subjected to intrinsic decomposition on Se-rich chain-like clusters and As-rich As2Se4/2 clusters based on homonuclear (As-As) bonding between the AsSe3/2 pyramids, acting as pre-cursors of As4Se4 thioarsenide molecular entities [61,62,63].
Thus, in a group of tetra-arsenic tetraselenide thioarsenides, the strong molecular-forming tendency becomes dominated with under-constrained (nc = 2.875) realgar-type As4Se4 cage molecules of D2d symmetry, with Ef ~ −72.713 kcal/mol, which is 0.40 kcal/mol better than Ef in the AsSe3/2 unit [23]. Among network derivatives from this As4Se4 molecule, the most plausible is an optimally constrained cluster (nc = 3.0) that appeared due to a single break in one of four Se atom positions (Ef ~ 0.25 kcal/mol), which can be labelled as x1-As4Se4 following the accepted CINCA nomenclature [60,61].
Among tetra-arsenic triselenide thioarsenides (corresponding to x = 57), the dimorphite-type As4Se3 cage molecule (under-constrained in view of nc = 2.71) composed of a triangular-pyramidal (As3)-As conformation is most plausible, possessing Ef ~ 0.33 kcal/mol (with respect to AsSe3/2 pyramid) [59]. All seven atoms in this molecule of C3v symmetry (isostructural with the molecule refined in α- and β-modifications of mineral dimorphite As4S3 by Whitfield [68,69]) are located in the same sphere, resulting in a specific 0D structure with low heat transfer and strong thermal expansion responses, which are accepted as characteristic features of the plastically crystalline As4Se3 modification [58,70]. Other molecular clusters of this type and their network-forming derivatives reconstructed by the break in Se atom positions are unfavorable, as compared with this dimorphite-type As4Se3 molecule [59], and, therefore, cannot be taken into consideration under comprehensive analysis of molecular–network disproportionality in arsenoselenides of similar chemical compositions.
In a group of tetra-arsenic biselenide (As4Se2) thioarsenides (x = 67), both molecular entities, the As4Se2-I cluster with nc = 2.67 composed of four (As-As) bonds in a zig-zag sequence (see Figure 8a) and the As4Se2-II cluster with nc = 2.50 composed of a (As-As) bond attached to the As3 triangle (Figure 8b), are unfavorable in view of Ef energy, approaching −4.42 kcal/mol and −3.59 kcal/mol, respectively. The same concerns network-forming clusters derived from these molecules by breaking in one of the Se atom positions, with both As4Se3H2 molecular prototypes reproduced in Figure 8c,d. But this situation cardinally changed in the transition to network clusters derived from these molecules by double breaking in both Se atom positions, with the H-saturated molecular prototypes of these clusters shown in Figure 8e,f. Surprisingly, the network-forming cluster (x2-As4Se2-I) derived from the As4Se2-I molecule, maintaining the closed tetragon-like As4 arrangement built of four (As-As) bonds (Figure 8e), has Ef ~ −0.72 kcal/mol, which is the best among all clusters in this compositional range [59]. Because of one small ring involved, the glassy network built of such clusters is rigid and stressed (nc = 3.33). These clusters facilitate the decomposition of dimorphite-type As4Se3 molecules into realgar-type As4Se4 ones, accompanied by the extraction of some amorphous products close to As4Se2 [59].
In a group of tetra-arsenic monoselenide (As4Se) thioarsenides (x = 80), where only two clusters are possible, the resolution is very simple. Any of these clusters, neither molecular As4Se composed of two edge-sharing As3 triangles (Figure 9a) nor network-forming x1-As4Se reproduced from this molecule by breaking in the Se atom position (see Figure 9b), can be stabilized in a realistic structure because of the very unfavorable cluster-forming energies Ef far beyond −5 kcal/mol.
Thus, in a group of tetra-arsenic As4 clusters (corresponding to x = 100 in AsxSe100−x system), the situation is non-trivial and complicated. The most stable amorphous and crystalline structures related to As polymorphs can be reconstructed from two principal building blocks typical for gaseous and condensed As states, these being, respectively, (i) the thioarsenide molecular cluster in the form of a regular four-atom pyramid-shaped As4 tetrahedron and (ii) the thioarsenide-type network-forming cluster in the form of a flattened pyramid-shaped unit derived through distortion from this regular As4-tetrahedral-like molecule by breaking in one of three As-As bonds at each As atom composed of a two-dimensional double-layer network of chair-configurated six-fold rings (As6⋅(2/3) = As4). These blocks within the arrangement of the three nearest neighbors can be differentiated by the Ef energies derived from the CINCA modelling, while distortion pathways beyond the three nearest neighbors explain the emergence of As allotropes in orthorhombic and rhombohedral structures [30], but this specificity cannot be accounted for within this cluster modelling routine. The under-constrained configuration of the tetrahedral As4 molecule (nc = 2.25) in Figure 10a is found to be very unfavorable in realistic structures in view of Ef ~ −4.31 kcal/mol. This finding is in line with the fact that yellow As (also named insulator, γ-As) consisting of As4 tetrahedra exists only in the gaseous state [71,72]. In contrast, the Ef energy of over-constrained network clusters (nc = 4.5) composed of a double-layer honeycomb structure of chair-configurated (As6⋅(2/3) = As4) rings (see Figure 10b) typical of both rhombohedral (grey or metallic, α-As) and orthorhombic (black or semiconducting, β-As) allotropes [30,31,71,72] seems very promising in view of Ef ~ −2.46 kcal/mol. This finding explains the appearance of the rhombohedral As phase as the most stable allotrope under crystallization in MQ-derived As-Se alloys.
Let us parameterize molecular–network disproportionality at the border of the glass-forming region in the As-Se system (close to As4Se2) based on the cluster-forming energies Ef calculated for the components of most expected decomposition scenarios, sketched in Figure 11.
The first scenario (Figure 11a) involves the decomposition of x2-As4Se2-I network-forming clusters (shown in Figure 8e) on dimorphite-type As4Se3 molecules [59] supplemented by the extraction of As6⋅(2/3) ring-like network-forming clusters (see Figure 10b), which occurs with a negative energetic barrier (ΔEf) corresponding to the spontaneous decomposition of the examined arsenoselenide alloy under MQ:
3⋅(x2-As4Se2-I) → 2 As4Se3 + As6⋅(2/3) + (ΔEf = −0.43 kcal/mol)
By removing down from As4Se2 stoichiometry, the second disproportionality scenario involving the spontaneous decomposition of x2-As4Se2-I network-forming clusters (Figure 8e) on realgar-type As4Se4 molecules [23] supplemented by the extraction of As6⋅(2/3) ring-like network clusters (Figure 10b) occurring also with a negative barrier becomes clear (see Figure 11b):
2⋅(x2-As4Se2-I) → As4Se4 + As6⋅(2/3) + (ΔEf = −0.17 kcal/mol)
Further approaching stoichiometric arsenic triselenide As2Se3, the third molecular–network disproportionality scenario, which involves the decomposition of x2-As4Se2-I network-forming clusters (Figure 8e) on layer-type As2Se3 network clusters composed of two corner-shared AsSe3/2 pyramids (as shown in [61,62]) supplemented by the extraction of As6⋅(2/3) ring-like network clusters (Figure 10b), occurring with a positive energetic barrier, can be validated (see Figure 11c):
3⋅(x2-As4Se2-I) → 2⋅As2Se3 + 2⋅As6⋅(2/3) + (ΔEf = +0.375 kcal/mol)
The positive value of the energetic barrier ΔEf in Equation (3) means that this decomposition scenario cannot be realized spontaneously but, rather, as activated thermal or milling-driven decomposition.
The spontaneous nature of molecular–network disproportionality in AsxSe100−x alloys from the border of the glass-forming region (65 < x < 70), obeying reactions (1) and (2), is proved by a large amount of molecular thioarsenides in these alloys stabilized under MQ with respect to the micro-RS spectra of unmilled samples (see Figure 4a,b and Figure 5a) and the crystalline inclusions of the rhombohedral As phase in these samples revealed by the XRPD analysis (see Figure 1, Figure 2 and Figure 3). Thus, from this disproportionality analysis, the compositional location of the border of the glass-forming region in the As-Se system nearby the As4Se2 thioarsenide composition (CN = 2.67) becomes completely understandable.
The nanomilling-driven molecular-to-network transition in As4Se2-bearing alloys is described by decomposition reaction (3), with a positive energetic barrier (ΔEf = 0.375 kcal/mol) comparable to the decomposition of As4Se3 molecules in As4Se4 ones and accompanying amorphous products (ΔEf = 0.41 kcal/mol) [59]. The transition towards more network conformations is accompanied by interplay between medium-range structure levels in As-Se alloys subjected to nanomilling (see Figure 1) and the respective broadening of their micro-RS spectra (see Figure 4c,d and Figure 5b). Since the remnants of thioarsenides destructed under grinding interact with oxygen, especially in PVP water solution, the cubic arsenolite As2O3 phase is intensively formed in these alloys subjected to nanomilling in combined dry–wet mode (see Figure 3). Because of the enhanced heterogeneity in the atomic-deficient microstructure, the superposition of the PAL spectra collected for PVP and MQ-derived umilled, dry-milled and dry–wet-milled samples shows a spectacular smoothly decaying trend over superimposed PAL spectra peaks and tails (highlighted in Figure 7).

3. Materials and Methods

3.1. Preparation, Preliminary Testing and Nanomilling of the Examined Arsenoselenides

The border of the glass-forming region in AsxSe100−x system from As-rich side is exemplified by MQ-derived alloys compositionally deviated around tetra-arsenic biselenide (As4Se2) thioarsenide stoichiometry, corresponding in coordination number CN = 2.67, these being under-stoichiometric glassy g-As65Se35 (CN = 2.65), stoichiometric glassy-crystalline g/c-As67Se33 (or g/c-As4Se2) and over-stochiometric glassy-crystalline g/c-As70Se30 (CN = 2.70). These alloys were synthesized via vibrational MQ from elemental precursors (As and Se of 5N purity), as described in more detail elsewhere [11,12,15,18]. The density ρ (±0.005 g⋅cm−3) of coarse-grained pieces of samples defined by the Archimedes displacement method in ethanol was lower in the glassy state (4.448 g⋅cm−3 for g-As65Se35) and higher in the glassy-crystalline state, approaching 4.654 g⋅cm−3 for g/c-As70Se30. The mean inter-atomic spacing dsm derived from the density of g/c-As67Se33 alloy was ~3.79 Å.
Mechanical treatment of the prepared alloys was performed in two stages.
Firstly, the MQ-derived ingots were subjected to nanostructurization by dry MM in a protective argon atmosphere in a planetary ball mill Pulverissete 6 (Fritsch, Idar-Oberstein, Germany). This grinding route was performed under a rotational speed of 500 rpm in 250 mL tungsten carbide chamber loaded with 50 tungsten carbide balls (each 10 mm in diameter) using 3 g of preliminary coarse-grained material sieved under 200 μm. Then, the fine-grained powder was pelletized in a stainless-steel die via ~0.7 GPa compressing, stabilizing dry-milled samples for further research.
In the second stage, part of the powder (prepared in a dry mode) was subjected to additional attrition in wet mode using a laboratory MiniCer mill (Netzsch, Germany), operational under a rotational speed of 3500 rpm. This 90-minute route was performed in 300 mL of 0.5% PVP water solution, milling shaft being loaded on 85% with yttrium-stabilized ZrO2 balls (each 0.6 mm in diameter). The PVP of ~40,000 g⋅mol−1 molecular weight purchased from Sigma-Aldrich Co. LLC (St. Louis, MO, USA) was used to prepare compositionally authentic nanosuspensions. Finally, they were dried at 70 °C and pelletized via compressing (under ~0.7 GPa) in stainless-steel die, stabilizing a set of the AsxSe100−x/PVP nanocomposites as plane-parallel disc-like pellets (~6 mm in diameter, ~1 mm in thickness), the most suitable shape for micro-RS and PAL measurements.
The particle size distribution was recognized for nanosuspensions of powdered As-Se alloys, employing photon cross-correlation spectroscopy with Nanophox particle size analyzer (Sympatec, Clausthal-Zellerfeld, Germany). The unimodal particle size distribution centered around ~180–190 nm was monitored in all cases. As an example, the particle size distribution in the suspension of nanomilled g-As65Se35 showing (x50)~182 nm (meaning that 50% of particles are smaller than 182 nm) and (x99)~291 nm (meaning that 99% of particles are smaller than 291 nm) is shown in Figure 12.

3.2. Atomic-Specific Phase Composition and Medium-Range Structure via Xrpd Analysis

The phase composition and medium-range structure of the arsenoselenides were recognized with the XRPD analysis using the STOE STADI P diffractometer operational in transmission mode of Cu Kα1-radiation (for details of the XRPD analysis, see [20,21,22,23,24]).
The phase diagram of As-Se alloys contains three stable crystalline compounds corresponding to different thioarsenide As4Sen stoichiometry, such as arsenic triselenide As2Se3 equivalent to tetra-arsenic hexaselenide As4Se6 (n = 6), arsenic monoselenide AsSe (tetra-arsenic tetraselenide As4Se4, n = 4) and tetra-arsenic triselenide As4Se3 (n = 3), existing as two normal crystalline polymorphs identified by Bastow and Whitfield [32] and nominated by Blachnik and Wickel [58] as ambient-temperature monoclinic α-As4Se3 and high-temperature orthorhombic α′-As4Se3. Under heating above 412 K, the α-As4Se3 transforms in α′-As4Se3 phase, and under further heating above 447 K, the latter transforms into plastically crystalline β-As4Se3 and unidentified amorphous phase, while only normal crystalline α′-As4Se3 phase can be obtained at ambient temperature by MQ [58]. Therefore, processing of the XRPD patterns was performed to recognize phase composition of the arsenoselenides using the data for As-Se polymorphs available from the known databases [73,74] and other resources, in part, the JCPDS card No. 65–2365 for monoclinic As2Se3 (space group s.g. P21/n, structure type s.t. α-As2S3, orpiment [35,36]), No. 73–0465 for trigonal Se (s.g. P3121 [37]), No. 71–0388 for monoclinic As4Se4 (s.g. P21/n, s.t. α-As4S4, realgar [33,34]), No. 04–4979 for orthorhombic As4Se3 (s.g. Pnma, s.t. α-As4S3, dimorphite [32]), No. 72–1048 for rhombohedral (grey) As or α-As (s.g. R 3 ¯ m [30,31]), No. 36–1490 for cubic arsenolite As2O3 (s.g. Fd3m [38]).
The amorphous phase in the examined alloys was identified due to diffuse peak halos in their XRPD patterning, in part, the FSDP (the first sharp diffraction peak), which is believed to be a signature of structural entities forming intermediate-range ordering in glass over a few tens Å reproduced in a reciprocal space near scattering vector Q1 ~(1–1.5) Ǻ−1 [75], and the SSDP (the second sharp diffraction peak, in terms of Elliott [76]) or PDP (the principal diffraction peak, in terms of Zeidler and Salmon [77]) serving as signature of extended-range order at Q2 ~(1.8–2.2)Ǻ−1. In the XRPD patterning of As-rich AsxSe100−x alloys (x > 40), the FSDP-related peak halo at ~(15–22)°2θ reflects correlations between some polyhedrons like thioarsenide As4Sen molecules, while the SSDP-related halo shifted to ~(28–33)°2θ reflects orientational specificity of these polyhedrons ascribed to the second-order pair atomic correlations close to the mean inter-atomic spacing, dsm [35]. At ~(50–60)°2θ (equivalent to Q3 ~3.3–4.0 Ǻ−1), the third diffraction peak (TDP) is observed as manifestation of shortest (~2.1–2.3 Ǻ) interatomic separation in a glass [35,77]. Thus, the XRPD measurement reveals three-peak structure of the patterns, which reflects a succession of single pairwise correlations defined by Q3 = QTDP and multi-pairwise correlations determined by Q1 = QFSDP and Q2 = QSSDP responsible for medium-range ordering [77].
Arrangement of diffuse peak halos in the XRPD patterns responsible for amorphous phase was analyzed using the STOE WinXPOW 3.03 [78] and PowderCell 2.4 [79] program packages, following normalization procedure with respect to the maximum. The error bar in the peak halo position (2θ) and full width at half maximum (FWHM) was not worse, ±0.05°2θ; the scattering vector position and width were calculated as Q = (4π/λ)⋅sinθ and ΔQ = (4π/λ)⋅sin(FWHM/2), respectively. The characteristic distance R (the spacing of peak halo responsible quasi-periodicity) and the correlation length L over which this quasi-periodicity were maintained and defined as for the Bragg diffraction (R = 2π/Q, L = 2π/ΔQ). Within modified microcrystalline approach [20,21,22,23,24], the diffuse peak-halo arrangement in the XRPD patterning in chalcogenide glass was also treated as arising from diffraction of coordination spheres, i.e., the shortest inter-atomic distances like in randomly packed multiparticulate systems (see [80] and literature therein), when the XRPD is governed by the Ehrenfest relation [81]:
2ds⋅sin θ = 1.23⋅λ
where ds is the average distance between scatterers (the radius of the coordination sphere). Noteworthily, the error bar in the above parameters (R, L, ds) does not exceed ±0.1 Å.

3.3. Atomic-Specific Microstructure by Micro-Rs Spectroscopy

The microstructure peculiarities of the arsenoselenides were also identified with RS microscopy using the Horiba Xplora apparatus equipped with CCD detector, all micro-RS spectra being collected at ambient temperature. The CW 785 nm laser of 90 mW output power was employed for excitation, the 10% power option being used to avoid photostructural effects. Other measurement options applied were as follows: x100 objective, 1800 1/mm grating, 500 μm hole and 50 μm slit. The spectral resolution was ~2 cm−1, and the spatial resolution was near ~2 μm. Number of scans was chosen depending on the surface of the samples to be sure that RS spectra processed with Horiba LabSpec 6 software were reasonably identical. The milled and unmilled samples were compared through normalization by matching the spectral areas of interest. The RS active bands were identified using the known data for chalcogenide compounds [3,41,42,43,44,45,46].

3.4. Atomic-Deficient Microstructure by Revised PAL Analysis

The PAL spectra were recorded with the fast-fast coincidence system ORTEC (230 ps in resolution) using 22Na isotope of ~50 kBq activity operated in normal measuring statistics (~1 M coincidences), as described in more detail elsewhere [29]. The best fitting of the PAL spectra was achieved with the LT 9.0 program [82] under decomposition into three negative exponentials obeying normalization (I1 + I2 + I3 = 1) and stabilizing the model-independent average positron lifetime τavΣ as the mass center of a whole PAL spectrum.
Under this measuring set-up, the reconstructed three-component PAL spectra allow for error bars in the lifetimes τi and intensities Ii at the level ±0.005 ns and 0.5%, respectively. This approach covers different channels in nanostructured substances arising from positrons annihilating in defect-free bulk states, intrinsic trapping sites (such as vacancy-type defects) and Ps-hosting holes (free-volume voids).
In nanostructured substances dominated by annihilation from positron- and Ps-related states, unconstrained three-component PAL spectra can be interpreted employing different mathematical algorithms based on canonical STM [28,29,57,83,84]. But this model is valid only for solids with one kind of trap having two-component PAL spectra. So, multichannel spectra of nanosubstances originating from a great variety of free-volume elements should be processed with STM modified for some restrictions. One of the most simplified is two-state STM ignoring Ps decaying. Within this approach, the positron-trapping modes (such as defect-specific τ2 and defect-free bulk τb lifetimes, positron trapping rate κd and a fraction of trapped positrons η) can be parameterized [57,83,84]. The most plausible free-volume void modification processes in nanostructured materials can be hypothesized as suggested by Shpotyuk M. et al. [84]. Finally, the ‘remainder’ over positron trapping in the PAL spectrum is ascribed to the Ps-decay channel owing to positrons annihilating as free particles or interacting with electrons from the environment [85]. The Ps localized in holes give indication on their radii R in terms of the longest τ3 lifetime with respect to the semiempirical Tao-Eldrup equation with material constants [85]. The fractional free volume (fv) of Ps-hosting sites can be calculated accepting the I3 intensity and some constants validated for materials without structural groups inhibiting Ps formation.
Alternatively, in substances obeying three-component PAL spectra with essential input of the third component, which cannot be recompensated by source contribution, two Ps-decay scenarios can be resolved in terms of fitting parameters only [56]. The prevailing trend in the Ps formation probability would result to iso-typical changes in both I1 and I3 intensities accompanied by opposite changes in I2 intensity. On the contrary, the prevailing trend in the hole density would result to iso-typical changes in both I2 and I3 intensities counterbalanced by opposite changes in I1 intensity.
In heterogeneous substances obeying selective localization of traps, positron annihilation occurs through mixed positron-trapping and Ps-decaying channels. The low-electron density holes with maximum free volume and minimum surface tension should fit to confine Ps stabilization owing to its repulsive exchange potential, while regions of higher negative electron density and polarization such as sub-nm voids are more suitable to capture electrically charged positrons [86]. So, positron annihilation is expected via interconnected positron-Ps-related channels, so that Ps-decay sites are the only free-volume holes which could be converted into positron traps (and vice versa). In this case, the generalized two-state STM modified for Ps-positron trapping conversion (known as the x3-x2-CDA, Coupling Decomposition Algorithm, or the Positronics approach [28,29]) can be used. Firstly, this algorithm is applied to an unstructured substance characterized by unconstrained three-component PAL spectrum, transforming it to generalized two-component form where all trapping inputs are gathered in the second component. Then, this procedure is performed for nanostructured substances (often modified by embedded nanoparticles, NP, thus defining respective nomenclature in parameterization). Nanostructurization-driven interplay between Ps- and positron-related sites can be refined by transition to differential spectrum with the first (τnIn) and second (τintIint) components defined under condition of full inter-channel balance in nanosubstance. Parameterization of the Ps-positron trapping conversion is performed within canonical two-state STM [28,29,57,83,84] applied to this differential PAL spectrum. Finally, trapping modes such as defect-specific τint and defect-free bulk τbNP positron lifetimes, trapping rate in defects κdNP, and some characteristics relevant to the size of these traps in terms of equivalent number of vacancies (τintτbNP) and nature of these traps in terms of (τintbNP), can be determined. Within this approach, directionality in trapping conversion (Ps-to-positron or positron-to-Ps) is defined by sign of both In and Iint intensities. So, this approach can be used as indicator separating meaningful nanostructurization-driven processes from simple interplay between uncorrelated positron-trapping and Ps-decaying channels.

3.5. Quantum–Chemical Modeling of Molecular–Network Clustering

The optimized configurations of thioarsenide As4Sen cage-like molecules (n = 6–0) and network-forming clusters derived from these molecules by breaking in available Se atom positions were reconstructed using ab initio quantum–chemical atomic cluster-modeling algorithm CINCA (Cation Interlinked Network Cluster Approach) [60,61]. The HyperChem Release 7.5 program based on restricted Hartree-Fock self-consistent field method with split-valence double-zeta basis set and single polarization function 6–311G* [87,88,89] was used. Optimization and single-point energy calculations were performed by the Fletcher–Reeves conjugate gradient method until a root-mean-square gradient of 0.1 kcal/(Å·mol) was reached. The cluster-forming energy (Ef) was corrected on the energy of terminated H atoms transforming network-forming cluster into molecular one [89,90] and recalculated with respect to the energy of a single AsSe3/2 pyramid (Ef = −72.309 kcal/mol [61]). This modelling route allows for the characterization of both molecular and network-forming conformations in covalent-bonded systems like chalcogenide glasses characterized by CN (coordination number), thus parameterizing the most energetically favorable molecular–network decomposition scenarios. To compare clusters accounting for small rings involved, the average number of constraints per atom nc was calculated following the Phillips–Thorpe constraint-counting algorithm with stretching and bending forces ascribed to intra-molecular bonds within the cluster [65,66,67].

4. Conclusions

Changes in the atomic-specific and atomic-deficient microstructure of binary AsxSe100−x arsenoselenides from the border of the glass-forming region (65 < x < 70) driven by nanomilling in dry and dry–wet modes are examined by multiexperimental characterization probes, including the X-ray powder diffraction (XRPD) in terms of the modified microcrystalline model, micro-Raman scattering (micro-RS) and revised positron annihilation lifetime (PAL) spectroscopy within the Positronics approach, complemented with molecular–network disproportionality analysis employing the ab initio quantum–chemical cluster-modeling algorithm CINCA (the Cation Interlinked Network Cluster Approach). The studied alloys are considered with respect to tetra-arsenic biselenide As4Se2 thioarsenide stoichiometry corresponding to coordination number CN = 2.67, these being glassy g-As65Se35 (CN = 2.65), glassy-crystalline g/c-As67Se33 (CN = 2.67) and g/c-As70Se30 (CN = 2.70).
Due to XRPD analysis, the crystalline inclusions in the examined alloys are ascribed to two phases enhanced under milling, especially in PVP (polyvinylpyrrolidone) water solution, such as rhombohedral As and cubic arsenolite As2O3. A straightforward interpretation of the observed changes in the amorphous structure of these alloys is developed by analyzing diffuse peak halos in their XRPD patterning recognizing the medium-range structure within the modified microcrystalline model. Disruption of the intermediate-range order due to weakening of the FSDP (the first sharp diffraction peak) responsible entities accompanied by an enhancement of the extended-range order due to fragmentation of the SSDP (the second sharp diffraction peak) responsible entities compose principal interplay between medium-range structure levels in arsenoselenides subjected to nanomilling. Since remnants of thioarsenide As4Sen molecular entities destructed under grinding interact with oxygen, a cubic arsenolite As2O3 phase is formed in the alloys subjected to nanomilling in wet mode.
Due to micro-RS spectroscopy, the examined alloys are stabilized by thioarsenide As4Sen molecules (preferentially with n = 3, 4) incorporated in the As-Se network. Nanomilling-driven molecular-to-network (amorphization) transitions occur due to the destruction of molecular thioarsenides followed by their incorporation in the network undergoing polyamorphic (reamorphization) transition. The examined arsenoselenides affected by generated defects become notably stressed, the respective defect-formation transformations being realized independently on crystalline inclusions in them.
Due to revised PAL analysis for direct conversion of bound positron–electron (Ps, positronium) states into positron traps, nanomilling-driven volumetric changes in PVP-stabilized alloys are identified with respect to dry-milled ones. Under wet milling, the Ps-hosting holes in a preferential PVP environment appear instead of positron traps with ~0.36–0.38 ns lifetimes corresponding to multiatomic vacancies in the As-Se matrix. Superposition of the PAL spectra collected for PVP, melt-quenching-derived, dry- and dry–wet-milled As-Se samples shows spectacular smoothly decaying behavior over superimposed PAL spectra peaks and tails.
The microstructure scenarios of spontaneous (under melt quenching) and activated (under nanomilling) decomposition of the As4Se2 network-forming clusters governing molecular–network disproportionality in AsxSe100−x alloys from the border of the glass-forming region (65 < x < 70) are recognized using the ab initio quantum–chemical cluster-modeling algorithm (CINCA). The over-constrained As6⋅(2/3)=As4 ring-like network clusters (acting as pre-cursors of rhombohedral As phase) are the main products of decomposition in As4Se2-bearing As-Se alloys. Two spontaneous processes ended by additional extraction of molecular thioarsenides with crystalline counterparts (such as dimorphite-type As4Se3 and realgar-type As4Se4) explain the location of the glass-forming border in the As-Se system near the As4Se2 composition (CN = 2.67), while the activated decomposition process ended by the extraction of layered As2Se3 network-forming entities composed of corner-shared AsSe3/2 pyramids is responsible for the nanomilling-driven molecular-to-network reamorphization transition in the examined alloys.

Author Contributions

Conceptualization, O.S. and P.B.; methodology, Y.S. and Z.L.B.; formal analysis, M.H., Y.S., P.D., A.I., V.B., R.W.-N., Z.L.B. and P.B.; investigation, M.H., Y.S., R.W.-N., Z.L.B., P.D., A.I. and V.B.; data curation, M.H., Y.S., Z.L.B., P.D., A.I. and V.B.; writing—original draft preparation, O.S.; writing—review and editing, Y.S. and P.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by the Slovak Research and Development Agency under contract SK-PL-23–0002 and the Scientific Grant Agency of the Ministry of Education, Science, Research and Sport of the Slovak Republic under contract 2/0112/22 (ZLB). This work is part of the research performed within project No. 0122U001806, the subject of the Program funded by the Ministry of Education and Science of Ukraine for the years 2022–2024 (OS). The project is co-financed by the Polish National Agency for Academic Exchange (agreement BPN/BUA/2021/1/00204/U/00001) and the Ministry of Education and Science of Ukraine in the frame of bilateral Ukrainian-Polish cooperation program for years 2023–2024. The part of research is performed within common Polish-Slovak research cooperation program for years 2024–2025 co-financed by Polish National Agency for Academic Exchange (agreement BPN/BSK/2023/1/00001/U/00001) and Slovak Research and Development Agency (project SK-PL-23-0002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. The raw data supporting the conclusions of this article will be made available by the authors on request.

Acknowledgments

Y.S. is grateful to SAIA for support within the National Scholarship Program of the Slovak Republic. This work was partly funded by the Simons Foundation, Award ID: 1290588 (P.D.).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Feltz, A.; Aust, H.; Blayer, A. Glass formation and properties of chalcogenide systems XXVI: Permittivity and the structure of glasses AsxSe1−x and GexSe1−x. J. Non-Cryst. Solids 1983, 55, 179–190. [Google Scholar] [CrossRef]
  2. Feltz, A. Amorphous Inorganic Materials and Glasses; VCH: Weinheim, Germany, 1993; pp. 1–446. [Google Scholar]
  3. Yang, G.; Bureau, B.; Rouxel, T.; Gueguen, Y.; Gulbiten, O.; Roiland, C.; Soignard, E.; Yarger, J.L.; Troles, J.; Sangleboeuf, J.-C.; et al. Correlation between structure and physical properties of chalcogenide glasses in the AsxSe1−x system. Phys. Rev. B 2010, 82, 195206. [Google Scholar] [CrossRef]
  4. Adam, J.-L.; Zhang, X. Chalcogenide Glasses: Preparation, Properties and Application; Woodhead Publishing Series in Electronic and Optical Materials: Philadelphia, PA, USA; New Delhi, India, 2013; pp. 1–717. [Google Scholar]
  5. Cui, S.; Chahal, R.; Boussard-Plédel, C.; Nazabal, V.; Doualan, J.-L.; Troles, J.; Lucas, J.; Bureau, B. From Selenium- to Tellurium-Based Glass Optical Fibers for Infrared Spectroscopies. Molecules 2013, 18, 5373–5388. [Google Scholar] [CrossRef]
  6. Cui, S.; Chahal, R.; Shpotyuk, Y.; Boussard, C.; Lucas, J.; Charpentier, F.; Tariel, H.; Loreal, O.; Nazabal, V.; Sire, O.; et al. Selenide and telluride glasses for mid-infrared bio-sensing. Proc. SPIE 2014, 8938, 893805. [Google Scholar] [CrossRef]
  7. Eggleton, B.J. Chalcogenide photonics: Fabrication, devices and applications Introduction. Opt. Express 2010, 18, 26632–26634. [Google Scholar] [CrossRef] [PubMed]
  8. Eggleton, B.; Luther-Davies, B.; Richardson, K. Chalcogenide photonics. Nat. Photon 2011, 5, 141–148. [Google Scholar] [CrossRef]
  9. Lucas, J.; Troles, J.; Zhang, X.H.; Boussard-Pledel, C.; Poulain, P.; Bureau, B. Glasses to see beyond visible. Des verres pour voir au-delà du visible. C.R. Chimie 2018, 21, 916–922. [Google Scholar] [CrossRef]
  10. Zhang, X.H.; Adam, J.L.; Bureau, B. Chalcogenide Glasses. In Springer Handbook of Glass. Springer Handbooks; Musgraves, J.D., Hu, J., Calvez, L., Eds.; Springer: Cham, Switzerland, 2019; pp. 525–552. [Google Scholar] [CrossRef]
  11. Mehta, N. Advances in Chalcogenide Glasses (ChGs): Past, Present, and Future Applications. In Advances in Glass Research. Advances in Material Research and Technology; Ikhmayies, S.J., Ed.; Springer: Cham, Switzerland, 2023; pp. 153–168. [Google Scholar] [CrossRef]
  12. Stenger, F.; Mende, S.; Schwedesc, J.; Peukertd, W. Nanomilling in stirred media mills. Chem. Eng. Sci. 2005, 60, 4557–4565. [Google Scholar] [CrossRef]
  13. Yadav, T.P.; Yadav, R.M.; Singh, D.P. Mechanical Milling: A Top Down Approach for the Synthesis of Nanomaterials and Nanocomposites. Nanosci. Nanotechnol. 2012, 2, 22–48. [Google Scholar] [CrossRef]
  14. Baláž, P.; Achimovicova, M.; Baláž, M.; Billik, P.; Cherkezova-Zheleva, Z.; Manuel Criado, J.; Delogu, F.; Dutkova, E.; Gaffet, E.; Gotor, F.J.; et al. Hallmarks of mechanochemistry: From nanoparticles to technology. Chem. Soc. Rev. 2013, 42, 7571–7637. [Google Scholar] [CrossRef]
  15. Baláž, P.; Baláž, M.; Achimovičová, M.; Bujňáková, Z.; Dutková, E. Chalcogenide mechanochemistry in materials science: Insight into synthesis and applications (a review). J. Mater. Sci. 2017, 52, 11851–11890. [Google Scholar] [CrossRef]
  16. El-Eskandarany, M.S.; Al-Hazza, A.; Al-Hajji, L.A.; Ali, N.; Al-Duweesh, A.A.; Banyan, M.; Al-Ajmi, F. Mechanical Milling: A Superior Nanotechnological Tool for Fabrication of Nanocrystalline and Nanocomposite Materials. Nanomaterials 2021, 11, 2484. [Google Scholar] [CrossRef]
  17. Senna, M. The Optimization of Mechanochemical Processes toward Functional Nanocomposite Materials. Powders 2023, 2, 659–677. [Google Scholar] [CrossRef]
  18. Bonazzi, P.; Bindi, L. A crystallographic review of arsenic sulfides: Effects of chemical variations and changes induced by exposure to light. Z. Kristallogr. 2008, 223, 132–147. [Google Scholar] [CrossRef]
  19. Gibbs, G.V.; Wallace, A.F.; Downs, R.T.; Ross, N.L.; Cox, D.F.; Rosso, K.M. Thioarsenides: A case for long-range Lewis acid–base-directed van der Waals interactions. Phys. Chem. Minerals 2010, 38, 267–291. [Google Scholar] [CrossRef]
  20. Shpotyuk, Y.; Boussard-Pledel, C.; Bureau, B.; Demchenko, P.; Szlezak, J.; Cebulski, J.; Bujňáková, Z.; Baláž, P.; Shpotyuk, O. Effect of high-energy mechanical milling on the FSDP-related XRPD correlations in Se-rich glassy arsenic selenides. J. Phys. Chem. Sol. 2019, 124, 318–326. [Google Scholar] [CrossRef]
  21. Shpotyuk, Y.; Demchenko, P.; Bujňáková, Z.; Baláž, P.; Boussard-Pledel, C.; Bureau, B.; Shpotyuk, O. Effect of high-energy mechanical milling on the medium-range ordering in glassy As-Se. J. Am. Ceram. Soc. 2020, 103, 1631–1646. [Google Scholar] [CrossRef]
  22. Shpotyuk, Y.; Shpotyuk, O.; Lukáčová Bujňáková, Z.; Baáž, P.; Hyla, M.; Boussard-Pledel, C.; Bureau, B. Tailoring Se-rich glassy arsenoselenides employing the nanomilling platform. Mater. Sci. Eng. B 2024, 300, 117069. [Google Scholar] [CrossRef]
  23. Shpotyuk, Y.; Demchenko, P.; Shpotyuk, O.; Balitska, V.; Boussard-Pledel, C.; Bureau, B.; Lukáčová Bujňáková, Z.; Baláž, P. High-energy mechanical milling-driven reamorphization in glassy arsenic monoselenide g-AsSe: On the path tailoring special molecular-network glasses. Materials 2021, 14, 4478. [Google Scholar] [CrossRef]
  24. Shpotyuk, O.; Ingram, A.; Bujnáková, Z.; Baláž, P.; Shpotyuk, Y.; Slobodzyan, D.; Wozny, M.; Boussard-Pledel, C.; Bureau, B. Free-Volume Structure of Polyvinylpyrrolidone-Capped Glassy As2Se3 Nanocomposites Prepared by Mechanical Milling. Polym. Eng. Sci. 2019, 59, 2438–2442. [Google Scholar] [CrossRef]
  25. Shpotyuk, O.; Ingram, A.; Filipecki, J.; Shpotyuk, Y.; Cebulski, J.; Lukáčová Bujňáková, Z.; Baláž, P. Volumetric nanostructurization in glassy arsenoselenides driven by high-energy mechanical dry- and wet-milling. Macromol. Sympos. 2022, 405, 2100253. [Google Scholar] [CrossRef]
  26. Shpotyuk, Y.; Ingram, A.; Shpotyuk, O.; Demchenko, P.; Lukáčová Bujňáková, Z.; Baláž, P. Polyvinylpyrrolidone-nanosized glassy arsenoselenides characterized by complementary Positronics and XRD analysis. Appl. Nanosci. 2023, 13, 4661–4674. [Google Scholar] [CrossRef]
  27. Kozdras, A.; Shpotyuk, O.; Mahlovanyi, B.; Shpotyuk, Y.; Kovalskiy, A. Thermodynamic heat-transfer phenomena in nanostructured glassy substances: A comparative study on g-As5Se95 and g-As55Se45. J. Therm. Anal. Calorim. 2023, 148, 2265–2271. [Google Scholar] [CrossRef]
  28. Shpotyuk, O.; Filipecki, J.; Ingram, A.; Golovchak, R.; Vakiv, M.; Klym, H.; Balitska, V.; Shpotyuk, M.; Kozdras, A. Positronics of subnanometer atomistic imperfections in solids as a high-informative structure characterization tool. Nanoscale Res. Lett. 2015, 10, 77. [Google Scholar] [CrossRef]
  29. Shpotyuk, O.; Ingram, A.; Boussard-Pledel, C.; Bureau, B.; Lukáčová Bujňáková, Z.; Baláž, P.; Mahlovanyi, B.; Shpotyuk, Y. The art of Positronics in contemporary nanomaterials science: A case study of sub-nanometer scaled glassy arsenoselenides. Materials 2022, 15, 302. [Google Scholar] [CrossRef] [PubMed]
  30. Greaves, G.N.; Elliott, S.R.; Davis, E.A. Amorphous arsenic. Adv. Phys. 1979, 28, 49–141. [Google Scholar] [CrossRef]
  31. Schiferl, D.; Barret, S. The Crystal Structure of Arsenic at 4.2, 78 and 299 °K. J. Appl. Cryst. 1969, 2, 30–36. [Google Scholar] [CrossRef]
  32. Bastow, T.J.; Whitfied, H.J. Crystal data and nuclear quadrupole resonance spectra of tetra-arsenic triselenide. J. Chem. Soc. Dalton Trans. 1977, 10, 959–961. [Google Scholar] [CrossRef]
  33. Bastow, T.J.; Whitfield, H.J.J. Crystal structure of tetra-arsenic tetraselenide. Chem. Soc. Dalton Trans 1973, 17, 1739–1740. [Google Scholar] [CrossRef]
  34. Smail, E.J.; Sheldrick, G.M. Tetra-arsenic tetraselenide. Acta Cryst. B 1973, 29, 2014–2016. [Google Scholar] [CrossRef]
  35. Renninger, A.L.; Averbach, B.L. Crystalline structures of As2Se3 and As4Se4. Acta Cryst. B 1973, 29, 1583–1589. [Google Scholar] [CrossRef]
  36. Stergiou, A.C.; Rentzeperis, P.J. The crystal structure of arsenic selenide, As2Se3. Zeitsch. Krist 1985, 173, 185–191. [Google Scholar] [CrossRef]
  37. Cherin, P.; Unger, P. The Crystal Structure of Trigonal Selenium. Inorg. Chem. 1967, 6, 1589–1591. [Google Scholar] [CrossRef]
  38. Ballirano, P.; Maras, A. Refinement of the crystal structure of arsenolite, As2O3. Ζ. Kristalogr. 2002, 217, 177–178. [Google Scholar] [CrossRef]
  39. Vaipolin, A.A.; Porai-Koshits, E.A. Structural models of glasses and the structures of crystalline chalcogenides. Sov. Phys. Solid State 1963, 5, 497–500. [Google Scholar]
  40. Tanaka, K.; Shimakawa, K. Amorphous Chalcogenide Semiconductors and Related Materials; Springer: New York, NY, USA; Dordrecht, The Netherlands; Heidelberg, Germany; London, UK, 2011; pp. 1–242. [Google Scholar]
  41. Lucovsky, G. Optic modes in amorphous As2S3 and As2Se3. Phys. Rev. B 1972, 6, 1480–1489. [Google Scholar] [CrossRef]
  42. Pangavhane, S.D.; Nemec, P.; Wagner, T.; Janca, J.; Havel, J. Laser desorption ionization time-of-flight mass spectrometric study of binary As-Se glasses. Rapid Commun. Mass Spectrom. 2010, 24, 2000–2008. [Google Scholar] [CrossRef] [PubMed]
  43. Lannin, J.S. Raman scattering properties of amorphous As and Sb. Phys. Rev. B 1977, 15, 3863–3871. [Google Scholar] [CrossRef]
  44. Brumbach, S.B.; Rosenblatt, G.M. In-cavity laser Raman spectroscopy of vapors at elevated temperatures. As4 and As4O6. J. Chem. Phys. 1972, 56, 3110–3117. [Google Scholar] [CrossRef]
  45. Kovanda, V.; Vlcek, M.; Jain, H. Structure of As-Se and As-P-Se glasses studied by Raman spectroscopy. J. Non-Cryst. Solids 2003, 326–327, 88–92. [Google Scholar] [CrossRef]
  46. Golovchak, R.; Oelgoetz, J.; Vlcek, M.; Esposito, A.; Saiter, A.; Saiter, J.-M.; Jain, H. Complex structural rearrangements in As-Se glasses. J. Chem. Phys. 2014, 140, 054505. [Google Scholar] [CrossRef]
  47. Kalyva, M.; Orava, J.; Siokou, A.; Pavlista, M.; Wagner, T.; Yannopoulos, S.N. Reversible Amorphous-to-Amorphous Transitions in Chalcogenide Films: Correlating Changes in Structure and Optical Properties. Adv. Funct. Mater. 2013, 23, 2052–2059. [Google Scholar] [CrossRef]
  48. Properzi, L.; Santoro, M.; Minicucci, M.; Iesari, F.; Ciambezi, M.; Nataf, L.; Le Godec, Y.; Irifune, T.; Baudelet, F.; Di Cicco, A. Structural evolution mechanisms of amorphous and liquid As2Se3 at high pressures. Phys. Rev. B 2016, 93, 214205. [Google Scholar] [CrossRef]
  49. Hu, Y.; Liang, J.; Xia, Y.; Zhao, C.; Jiang, M.H.; Ma, J.; Tie, Z.X.; Jin, Z. 2D Arsenene and Arsenic Materials: Fundamental Properties, Preparation, and Applications. Small 2021, 18, 2104556. [Google Scholar] [CrossRef]
  50. Shpotyuk, O.; Bujňáková, Z.; Baláž, P.; Ingram, A.; Shpotyuk, Y. Positron annihilation lifetime study of polyvinylpyrrolidone for nanoparticle-stabilizing pharmaceuticals. J. Pharm. Biomed. Anal. 2016, 117, 419–425. [Google Scholar] [CrossRef]
  51. Jensen, K.O.; Salmon, P.S.; Penfold, I.T.; Coleman, P.G. Microvoids in chalcogenide glasses studied by positron annihilation. J. Non-Cryst. Solids 1994, 170, 57–64. [Google Scholar] [CrossRef]
  52. Ingram, A.; Golovchak, R.; Kostrzewa, M.; Wacke, S.; Shpotyuk, M.; Shpotyuk, O. Compositional dependences of average positron lifetime in binary As-S/Se glasses. Phys. B Condens. Matter 2012, 407, 652–655. [Google Scholar] [CrossRef]
  53. Shpotyuk, O.; Golovchak, R.; Ingram, A.; Boyko, V.; Shpotyuk, L. Comparative study of extended free-volume defects in As- and Ge-based glassy semiconductors: Theoretical prediction and experimental probing with PAL technique. Phys. Stat. Solidi C 2013, 10, 117–120. [Google Scholar] [CrossRef]
  54. Shpotyuk, O.; Ingram, A.; Shpotyuk, M.; Filipecki, J. Prediction of free-volume-type correlations in glassy chalcogenides from positron lifetime measurements. Nucl. Instr. Meth. Phys. Res. B 2014, 338, 66–71. [Google Scholar] [CrossRef]
  55. Shpotyuk, O.; Filipecki, J.; Shpotyuk, M.; Ingram, A. Free volume evolution in chalcogenide glasses as probed by PAL spectroscopy. Solid State Ionics 2014, 267, 38–43. [Google Scholar] [CrossRef]
  56. Vijay, Y.K.; Wate, S.; Awasthi, D.K.; Das, D.; Ghughre, S. Ion induced effects in polymers. Indian J. Eng. Mater. Sci. 2000, 7, 375–377. [Google Scholar]
  57. Krause-Rehberg, R.; Leipner, H. Positron Annihilation in Semiconductors: Defect Studies; Springer: Heidelberg, Germany, 1999; pp. 1–350. [Google Scholar]
  58. Blachnik, R.; Wickel, U. Thermal behaviour of A4B3 cage molecules (A = P, As; B = S, Se). Thermochim. Acta 1984, 81, 185–196. [Google Scholar] [CrossRef]
  59. Shpotyuk, O.; Hyla, M.; Shpotyuk, Y.; Lukáčová Bujňáková, Z.; Baláž, P.; Demchenko, P.; Kozdraś, A.; Boyko, V.; Kovalskiy, A. Molecular-Network Transformations in Tetra-Arsenic Triselenide Glassy Alloys Tuned within Nanomilling Platform. Molecules 2024, 29, 3245. [Google Scholar] [CrossRef] [PubMed]
  60. Shpotyuk, O.; Hyla, M.; Boyko, V. Structural-topological genesis of network-forming nano-clusters in chalcogenide semiconductor glasses. J. Optoelectron. Adv. Mater. 2013, 15, 1429–1437. [Google Scholar]
  61. Shpotyuk, O.; Hyla, M.; Boyko, V. Compositionally-dependent structural variations in glassy chalcogenides: The case of binary As-Se system. Comput. Mater. Sci. 2015, 110, 144–151. [Google Scholar] [CrossRef]
  62. Hyla, M.; Shpotyuk, O. Topological configurations of principal network-forming clusters in As/Ge-Se glasses. J. Optoelectron. Adv. Mater. 2017, 19, 211–217. [Google Scholar]
  63. Hyla, M. Compositional anomalies in Se-rich As-Se glass-forming system as revealed from positron annihilation lifetime spectroscopy and ab initio cluster modeling. Comput. Mater. Sci. 2017, 135, 165–168. [Google Scholar] [CrossRef]
  64. Hyla, M. Network-Forming Nanoclusters in Binary As–S/Se Glasses: From Ab Initio Quantum Chemical Modeling to Experimental Evidences. Nanoscale Res. Lett. 2017, 12, 45. [Google Scholar] [CrossRef] [PubMed]
  65. Phillips, J.C. Topology of covalent non-crystalline solids. I: Short-range order in chalcogenide alloys. J. Non-Cryst. Solids 1979, 34, 153–181. [Google Scholar] [CrossRef]
  66. Thorpe, M.F. Continuous deformations in random networks. J. Non-Cryst. Solids 1983, 57, 355–370. [Google Scholar] [CrossRef]
  67. Thorpe, M.F. Bulk and surface floppy modes. J. Non-Cryst. Solids 1995, 182, 135–142. [Google Scholar] [CrossRef]
  68. Whitfield, H.J. The crystal structure of tetra-arsenic trisulphide. J. Chem. Soc. A 1970, 1800–1803. [Google Scholar] [CrossRef]
  69. Whitfield, H.J. Crystal structure of the β-form of tetra-arsenic trisulphide. J. Chem. Soc. Dalton 1973, 17, 1737–1738. [Google Scholar] [CrossRef]
  70. Blachnik, R.; Hoppe, A.; Wickel, U. Die Systeme Arsen-Schwefel und Arsen-Selen und die thermodynamischen Daten ihrer Verbindungen. Z. Anorg. Allg. Chem. 1980, 468, 78–90. [Google Scholar] [CrossRef]
  71. Seidl, M.; Balazs, G.; Scgeer, M. The Chemistry of Yellow Arsenic. Chem. Revs. 2019, 119, 8406–8434. [Google Scholar] [CrossRef] [PubMed]
  72. Hart, M.; Chen, J.; Michaelides, A.; Sella, A.; Shaffer, M.S.P.; Salzmann, C.G. One-dimensional Arsenic Allotropes: Polymerization of Yellow Arsenic inside Single-wall Carbon Nanotubes. Angew. Chem. 2018, 130, 11823–11827. [Google Scholar] [CrossRef]
  73. Downs, R.T.; Hall-Wallace, M. The American mineralogist crystal structure database. Am. Mineral. 2003, 88, 247–250. [Google Scholar]
  74. Villars, P.; Cenzual, K. Pearson’s Crystal Data: Crystal Structure Database for Inorganic Compounds; Release 2014/15; ASM International: Materials Park, OH, USA, 2014. [Google Scholar]
  75. Elliott, S.R. Extended-range order, interstitial voids and the first sharp diffraction peak of network glasses. J. Non-Cryst. Solids 1995, 182, 40–48. [Google Scholar] [CrossRef]
  76. Elliott, S.R. Second sharp diffraction peak in the structure factor of binary covalent network glasses. Phys. Rev. B 1995, 51, 8599–8601. [Google Scholar] [CrossRef]
  77. Zeidler, A.; Salmon, P.S. Pressure-driven transformation of the ordering in amorphous network-forming materials. Phys. Rev. B 2016, 93, 214204. [Google Scholar] [CrossRef]
  78. Roisnel, T.; Rodriguez-Carvajal, J. WinPLOTR: A Windows tool for powder diffraction pattern analysis. Mater. Sci. Forum 2001, 118, 378–381. [Google Scholar] [CrossRef]
  79. Kraus, W.; Nolze, G. POWDER CELL—A program for the representation and manipulation of crystal structures and calculation of the resulting X-ray powder patterns. J. Appl. Cryst. 1996, 29, 301–303. [Google Scholar] [CrossRef]
  80. Feng, R.; Stachurski, Z.H.; Rodrigues, M.D.; Kluth, P.; Araujo, L.L.; Bulla, D.; Ridway, M.C. X-ray scattering from amorphous solids. J. Non-Cryst. Solids 2013, 383, 21–27. [Google Scholar] [CrossRef]
  81. Ehrenfest, P. On interference phenomena to be expected when Roentgen rays pass through a diatomic gas. Proc. KNAW 1915, 17, 1184–1190. [Google Scholar]
  82. Kansy, J. Microcomputer program for analysis of positron annihilation lifetime spectra. Nucl. Instr. Meth. Phys. Res. A 1996, 374, 235–244. [Google Scholar] [CrossRef]
  83. Tuomisto, F.; Makkonen, I. Defect identification in semiconductors with positron annihilation: Experiment and theory. Rev. Mod. Phys. 2013, 85, 1583–1631. [Google Scholar] [CrossRef]
  84. Shpotyuk, M.; Ingram, A.; Shpotyuk, O. Characterization of radiation-induced effects in amorphous arsenic sulfides by positron annihilation lifetime spectroscopy. J. Mater. Res. 2015, 30, 1422–1429. [Google Scholar] [CrossRef]
  85. Jean, Y.C.; Mallon, P.E.; Schrader, D.M. Principles and Application of Positron and Positronium Chemistry; World Scientific Publishing Co Pte Ltd.: Hackensack, NJ, USA; London, UK; Singapore; Hong Kong, China, 2003; pp. 1–406. [Google Scholar]
  86. Castelli, F.; Consolatti, G.; Marlotti, G.T. Positronium confined in nanocavities: The role of electron exchange correlations. Nanomaterials 2021, 11, 2350–2351. [Google Scholar] [CrossRef] [PubMed]
  87. Hehre, W.J.; Stewart, R.F.; Pople, J.A. Self-consistent molecular-orbital methods. I. Use of Gaussian expansions of slater-type atomic orbitals. J. Chem. Phys. 1969, 51, 2657–2665. [Google Scholar] [CrossRef]
  88. McLean, A.D.; Chandler, G.S. Contracted Gaussian basis sets for molecular calculations. I. Second row atoms, Z = 11–18. J. Chem. Phys. 1980, 72, 5639–5648. [Google Scholar] [CrossRef]
  89. Jackson, K. Electric fields in electronic structure calculations: Electric polarizabilities and IR and Raman spectra from first principles. Phys. Stat. Solidi B 2000, 217, 293–310. [Google Scholar] [CrossRef]
  90. Holomb, R.; Veres, M.; Mitsa, V. Ring-, branchy-, and cage-like AsnSm nanoclusters in the structure of amorphous semiconductors: Ab initio and Raman study. J. Optoelectron. Adv. Mater. 2009, 11, 917–923. [Google Scholar]
Figure 1. The normalized XRPD patterns of MQ-derived g/c-As67Se33 in unmilled and dry-milled state showing three principle diffuse peak halos responsible for the FSDP (~15–25°2θ), SSDP (~28–33°2θ) and TDP (~50–60°2θ). The Bragg-diffraction reflexes of crystalline counterparts are reproduced (from the top to the bottom) in a sequence: rhombohedral (grey) As (JCPDS No. 72–1048) [30,31], orthorhombic As4Se3 (JCPDS No. 04–4979) [32], monoclinic As4Se4 (JCPDS No. 71–0388) [33,34], monoclinic As2Se3 (JCPDS No. 65–2365) [35,36] and trigonal Se (JCPDS No. 73–0465) [37].
Figure 1. The normalized XRPD patterns of MQ-derived g/c-As67Se33 in unmilled and dry-milled state showing three principle diffuse peak halos responsible for the FSDP (~15–25°2θ), SSDP (~28–33°2θ) and TDP (~50–60°2θ). The Bragg-diffraction reflexes of crystalline counterparts are reproduced (from the top to the bottom) in a sequence: rhombohedral (grey) As (JCPDS No. 72–1048) [30,31], orthorhombic As4Se3 (JCPDS No. 04–4979) [32], monoclinic As4Se4 (JCPDS No. 71–0388) [33,34], monoclinic As2Se3 (JCPDS No. 65–2365) [35,36] and trigonal Se (JCPDS No. 73–0465) [37].
Molecules 29 03948 g001
Figure 2. The normalized XRPD patterns of MQ-derived g/c-As67Se33 before (a) and after nanomilling in dry (b) and combined dry–wet (c) mode showing three principal diffuse peak halos in comparison with the Bragg-diffraction reflexes from rhombohedral As (JCPDS No. 72–1048) [30,31] and cubic arsenolite As2O3 phase (JCPDS No. 36–1490) [38].
Figure 2. The normalized XRPD patterns of MQ-derived g/c-As67Se33 before (a) and after nanomilling in dry (b) and combined dry–wet (c) mode showing three principal diffuse peak halos in comparison with the Bragg-diffraction reflexes from rhombohedral As (JCPDS No. 72–1048) [30,31] and cubic arsenolite As2O3 phase (JCPDS No. 36–1490) [38].
Molecules 29 03948 g002
Figure 3. The normalized XRPD patterns of g-As65Se35 (black curve) and g/c-As70Se30 (red curve) after nanomilling in dry–wet mode, showing three principal diffuse peak halos corresponding to ‘amorphous’ phase overlapped with sharp broadened Bragg-diffraction reflexes originated from the planes (111) at 13.86°2θ (d = 6.390Å, I = 63%), (222) at 27.90°2θ (d = 3.195Å, I = 100%), (400) at 32.33°2θ (d = 2.769Å, I = 27%), (331) at 35.32°2θ (d = 2.541 Å, I = 38%), (440) at 46.36°2θ (d = 1.957Å, I = 27%) and (551) at 59.59°2θ (d = 1.551 Å, I = 27%) in cubic structure of arsenolite As2O3 (JCPDS No. 36–1490) [38].
Figure 3. The normalized XRPD patterns of g-As65Se35 (black curve) and g/c-As70Se30 (red curve) after nanomilling in dry–wet mode, showing three principal diffuse peak halos corresponding to ‘amorphous’ phase overlapped with sharp broadened Bragg-diffraction reflexes originated from the planes (111) at 13.86°2θ (d = 6.390Å, I = 63%), (222) at 27.90°2θ (d = 3.195Å, I = 100%), (400) at 32.33°2θ (d = 2.769Å, I = 27%), (331) at 35.32°2θ (d = 2.541 Å, I = 38%), (440) at 46.36°2θ (d = 1.957Å, I = 27%) and (551) at 59.59°2θ (d = 1.551 Å, I = 27%) in cubic structure of arsenolite As2O3 (JCPDS No. 36–1490) [38].
Molecules 29 03948 g003
Figure 4. The normalized micro-RS spectra of MQ-derived g/c-As67Se33 reproduced in a sequence from the bottom to the top: (a) unmilled bulk pieces (unpelletized); (b) pelletized coarse-grained sample; (c) pelletized dry-milled sample; (d) pelletized dry–wet-milled sample. The most prominent features in the micro-RS spectrum of the bulk unmilled sample (a) are marked by vertical arrows, and traced by dotted lines to the respective micro-RS spectra of the pelletized samples (bd).
Figure 4. The normalized micro-RS spectra of MQ-derived g/c-As67Se33 reproduced in a sequence from the bottom to the top: (a) unmilled bulk pieces (unpelletized); (b) pelletized coarse-grained sample; (c) pelletized dry-milled sample; (d) pelletized dry–wet-milled sample. The most prominent features in the micro-RS spectrum of the bulk unmilled sample (a) are marked by vertical arrows, and traced by dotted lines to the respective micro-RS spectra of the pelletized samples (bd).
Molecules 29 03948 g004
Figure 5. The normalized micro-RS spectra of MQ-derived g-As65Se35 in unmilled state (a) and after nanomilling in a single dry mode (b). The most prominent features in the micro-RS spectrum of the bulk sample (a) are marked by arrows and traced by dotted lines to that of dry-milled sample (b).
Figure 5. The normalized micro-RS spectra of MQ-derived g-As65Se35 in unmilled state (a) and after nanomilling in a single dry mode (b). The most prominent features in the micro-RS spectrum of the bulk sample (a) are marked by arrows and traced by dotted lines to that of dry-milled sample (b).
Molecules 29 03948 g005
Figure 6. The raw PAL spectra of pelletized g/c-As70Se30 in unmilled state (a) and after nanomilling in dry mode (b) and combined dry–wet mode (c) as compared with the spectrum of PVP pelletized under the same conditions (d). The collected PAL spectra are reconstructed from unconstrained three-term fitting and reproduced at background of source contribution with bottom insets showing statistical scatter of variance. The occupation of “tail” states in unmilled and dry-milled samples grows notably under transition to dry–wet-milled sample approaching that in the pelletized PVP.
Figure 6. The raw PAL spectra of pelletized g/c-As70Se30 in unmilled state (a) and after nanomilling in dry mode (b) and combined dry–wet mode (c) as compared with the spectrum of PVP pelletized under the same conditions (d). The collected PAL spectra are reconstructed from unconstrained three-term fitting and reproduced at background of source contribution with bottom insets showing statistical scatter of variance. The occupation of “tail” states in unmilled and dry-milled samples grows notably under transition to dry–wet-milled sample approaching that in the pelletized PVP.
Molecules 29 03948 g006
Figure 7. The overlapping of the PAL spectra in the examined arsenoselenides g-As65Se35 (a), g/c-As67Se33 (b) and g/c-As70Se30 (c) pelletized before nanomilling (black points) and after nanomilling in dry mode (red points) and dry–wet mode (green points) as compared with the PAL spectrum in the PVP sample pelletized under the same conditions (blue points). The insets show a nearly invariant tendency in the PAL spectra peaks depressed in the right wing after nanomilling in dry–wet mode due to moderated Ps-formation probability and slightly changed average positron lifetime τav. The changes in the PAL spectra tails of unmilled, dry- and dry–wet-milled samples are due to increase in density of o-Ps hosting holes. There is no evident empty gap between the PAL spectra tails for dry-milled and dry–wet-milled glassy samples as compared with glassy-crystalline samples caused by changes in Ps decaying states under transition to annihilation in PVP-bearing medium.
Figure 7. The overlapping of the PAL spectra in the examined arsenoselenides g-As65Se35 (a), g/c-As67Se33 (b) and g/c-As70Se30 (c) pelletized before nanomilling (black points) and after nanomilling in dry mode (red points) and dry–wet mode (green points) as compared with the PAL spectrum in the PVP sample pelletized under the same conditions (blue points). The insets show a nearly invariant tendency in the PAL spectra peaks depressed in the right wing after nanomilling in dry–wet mode due to moderated Ps-formation probability and slightly changed average positron lifetime τav. The changes in the PAL spectra tails of unmilled, dry- and dry–wet-milled samples are due to increase in density of o-Ps hosting holes. There is no evident empty gap between the PAL spectra tails for dry-milled and dry–wet-milled glassy samples as compared with glassy-crystalline samples caused by changes in Ps decaying states under transition to annihilation in PVP-bearing medium.
Molecules 29 03948 g007
Figure 8. The ball-and-stick presentation of optimized configuration of tetra-arsenic biselenide thioarsenide As4Se2-I molecule composed by four (As-As) bonds in zig-zag sequence (a) and As4Se2-II molecule composed by (As-As) bond attached to As3 triangle (b), as compared with As4Se3H2 and As4Se4H4 molecular prototypes of network clusters derived from these molecules by single (x1-As4Se2-I—(c), x1-As4Se2-II—(d)) and double (x2-As4Se2-I—(e), x2-As4Se2-II—(f)) breaking in available Se atom positions. The cluster-forming energies Ef are given in respect to AsSe3/2 pyramid (Ef = −72.309 kcal/mol [60]). The H, Se and As atoms are, respectively, grey-, blue- and red-colored, and chemical bonds between atoms are denoted by respectively colored sticks. The average number of constraints nc is given following the Phillips-Thorpe constraint-counting algorithm [65,66,67].
Figure 8. The ball-and-stick presentation of optimized configuration of tetra-arsenic biselenide thioarsenide As4Se2-I molecule composed by four (As-As) bonds in zig-zag sequence (a) and As4Se2-II molecule composed by (As-As) bond attached to As3 triangle (b), as compared with As4Se3H2 and As4Se4H4 molecular prototypes of network clusters derived from these molecules by single (x1-As4Se2-I—(c), x1-As4Se2-II—(d)) and double (x2-As4Se2-I—(e), x2-As4Se2-II—(f)) breaking in available Se atom positions. The cluster-forming energies Ef are given in respect to AsSe3/2 pyramid (Ef = −72.309 kcal/mol [60]). The H, Se and As atoms are, respectively, grey-, blue- and red-colored, and chemical bonds between atoms are denoted by respectively colored sticks. The average number of constraints nc is given following the Phillips-Thorpe constraint-counting algorithm [65,66,67].
Molecules 29 03948 g008
Figure 9. The ball-and-stick presentation of optimized configuration of tetra-arsenic monoselenide As4Se molecule composed by two edge-sharing As3 triangles (a), as compared with As4Se2H2 molecular prototype of network-forming cluster derived from this molecule by breaking in Se atom position x1-As4Se (b). The terminated H atoms are grey-colored, Se and As atoms are respectively blue- and red-colored, chemical bonds between atoms are denoted by, respectively, colored sticks.
Figure 9. The ball-and-stick presentation of optimized configuration of tetra-arsenic monoselenide As4Se molecule composed by two edge-sharing As3 triangles (a), as compared with As4Se2H2 molecular prototype of network-forming cluster derived from this molecule by breaking in Se atom position x1-As4Se (b). The terminated H atoms are grey-colored, Se and As atoms are respectively blue- and red-colored, chemical bonds between atoms are denoted by, respectively, colored sticks.
Molecules 29 03948 g009
Figure 10. The ball-and-stick presentation of optimized configuration of regular pyramid-shaped tetra-arsenic As4 thioarsenide molecule (a) and As6H6 molecular prototype of network-forming cluster derived by distortion from this molecule in a form of flattened pyramidal-shaped unit, which composes two-dimensional double-layer network of chair-configurated six-fold rings of As atoms (As6⋅(2/3) = As4). The H and As atoms are respectively grey- and red-colored, and chemical bonds between atoms are denoted by respectively colored sticks.
Figure 10. The ball-and-stick presentation of optimized configuration of regular pyramid-shaped tetra-arsenic As4 thioarsenide molecule (a) and As6H6 molecular prototype of network-forming cluster derived by distortion from this molecule in a form of flattened pyramidal-shaped unit, which composes two-dimensional double-layer network of chair-configurated six-fold rings of As atoms (As6⋅(2/3) = As4). The H and As atoms are respectively grey- and red-colored, and chemical bonds between atoms are denoted by respectively colored sticks.
Molecules 29 03948 g010
Figure 11. Three decomposition scenarios of x2-As4Se2-I network-forming clusters governing molecular-network disproportionality in tetra-arsenic biselenide As4Se2-bearing arsenoselenides: (a)—spontaneous decomposition under ΔEf = –0.43 kcal/mol; (b)—spontaneous decomposition under ΔEf = –0.17 kcal/mol; (c)—activated decomposition under ΔEf = +0.375 kcal/mol. The terminated H atoms are grey-colored, Se and As atoms are, respectively, blue- and red-colored, and chemical bonds between atoms are denoted by the respective colored sticks.
Figure 11. Three decomposition scenarios of x2-As4Se2-I network-forming clusters governing molecular-network disproportionality in tetra-arsenic biselenide As4Se2-bearing arsenoselenides: (a)—spontaneous decomposition under ΔEf = –0.43 kcal/mol; (b)—spontaneous decomposition under ΔEf = –0.17 kcal/mol; (c)—activated decomposition under ΔEf = +0.375 kcal/mol. The terminated H atoms are grey-colored, Se and As atoms are, respectively, blue- and red-colored, and chemical bonds between atoms are denoted by the respective colored sticks.
Molecules 29 03948 g011
Figure 12. The unimodal particle size distribution in nanosuspension of MQ-derived g-As65Se35 showing the parameters (x50) and (x99), respectively, approaching ~182 nm and ~291 nm.
Figure 12. The unimodal particle size distribution in nanosuspension of MQ-derived g-As65Se35 showing the parameters (x50) and (x99), respectively, approaching ~182 nm and ~291 nm.
Molecules 29 03948 g012
Table 1. Parameterization of the FSDP-related diffuse peak halo in the examined MQ-derived arsenoselenides before and after nanomilling in dry and combined dry–wet mode.
Table 1. Parameterization of the FSDP-related diffuse peak halo in the examined MQ-derived arsenoselenides before and after nanomilling in dry and combined dry–wet mode.
AsxSe100−x Samples:2θFWHMQ1ΔQ1RLds
Phase Composition, CNState°2θ°2θÅ−1Å−1ÅÅÅ
g-As65Se35, CN = 2.65unmilled15.641(5)2.49(1)1.110.185.6635.526.96
dry-milled16.170(13)4.42(4)1.150.315.4819.986.74
dry–wet-milled15.627(42)6.20(12)1.110.445.6714.256.97
g/c-As67Se33, CN = 2.67unmilled15.479(4)2.34(1)1.100.175.7137.87.02
dry-milled15.732(10)3.77(6)1.120.275.6323.46.92
dry–wet-milled15.891(15)5.08(4)1.130.365.5717.46.85
g/c-As70Se30, CN = 2.70unmilled15.318(5)2.33(1)1.090.175.7837.907.11
dry-milled15.668(4)4.38(5)1.110.315.6520.176.95
dry–wet-milled15.605(5)8.33(12)1.110.595.6710.656.97
Table 2. The best-fit PAL spectra parameterization in the pelletized arsenoselenide samples in unmilled state and after nanomilling in dry- and dry–wet modes compared with these parameters in the PVP samples pelletized under the same conditions [50].
Table 2. The best-fit PAL spectra parameterization in the pelletized arsenoselenide samples in unmilled state and after nanomilling in dry- and dry–wet modes compared with these parameters in the PVP samples pelletized under the same conditions [50].
Pelletized AsxSe100−x Samples:
Phase Composition, State
[FIT-1]τ1,
ns
τ2,
ns
τ3,
ns
I2,
a.u.
I3,
a.u.
τav.,
ns
PVP pelletized [50]0.2000.1960.4721.8670.2560.1190.466
g-As65Se35, unmilled0.0600.2070.3672.0890.5200.0150.318
g-As65Se35, dry-milled0.0400.1980.3712.1800.5600.0150.325
g-As65Se35, dry–wet-milled0.0700.2160.4311.9370.3900.0460.379
g/c-As67Se33, unmilled0.0900.2010.3622.0650.5500.0160.318
g/c-As67Se33, dry-milled0.0500.1880.3622.0950.6000.0150.323
g/c-As67Se33, dry–wet-milled0.0700.2020.4031.9140.4700.0490.379
g/c-As70Se30, unmilled0.0480.2180.3862.0190.3200.0110.291
g/c-As70Se30, dry-milled0.0440.2180.3852.1200.3900.0130.307
g/c-As70Se30, dry–wet-milled0.0700.2120.4251.7300.3290.0620.377
Table 3. The positron trapping modes in the pelletized arsenoselenides in unmilled state and after nanomilling in dry and dry–wet modes compared with these parameters in the PVP samples pelletized under the same conditions [50], derived from their reconstructed three-component PAL spectra treated within two-state simple-trapping model (STM) ignoring Ps-decaying contribution.
Table 3. The positron trapping modes in the pelletized arsenoselenides in unmilled state and after nanomilling in dry and dry–wet modes compared with these parameters in the PVP samples pelletized under the same conditions [50], derived from their reconstructed three-component PAL spectra treated within two-state simple-trapping model (STM) ignoring Ps-decaying contribution.
Pelletized AsxSe100−x Samples: Phase Composition, StatePositron-Trapping ModesPs-Decay Modes
τb,
ns
κd,
ns−1
τ2−τb,
ns
τ2b,
a.u.
η,
a.u.
R3,
nm
fv3,
%
PVP pelletized [50]0.240.870.241.990.170.2761.88
g-As65Se35, unmilled0.271.110.101.370.230.2960.30
g-As65Se35, dry-milled0.271.350.101.380.270.3040.32
g-As65Se35, dry–wet-milled0.270.950.161.590.210.2820.78
g/c-As67Se33, unmilled0.271.220.091.350.240.2930.30
g/c-As67Se33, dry-milled0.271.560.101.360.290.2970.30
g/c-As67Se33, dry–wet-milled0.271.220.141.510.250.2790.78
g/c-As70Se30, unmilled0.250.640.131.520.210.2890.14
g/c-As70Se30, dry-milled0.260.780.121.470.170.2980.25
g/c-As70Se30, dry–wet-milled0.2670.830.171.650.170.2610.84
Table 4. Trapping conversion modes calculated for pelletized dry-milled arsenoselenide sample respectively to the same sample subjected to nanomilling in combined dry–wet mode.
Table 4. Trapping conversion modes calculated for pelletized dry-milled arsenoselenide sample respectively to the same sample subjected to nanomilling in combined dry–wet mode.
Pelletized AsxSe100−x SampleI ComponentII ComponentτavNPTrapping-Conversion Modes
τnInτintIintτbNPκdNPτintτbNPτint/τbNP
nsa.u.nsa.u.nsnsns−1nsa.u.
g-As65Se350.1930.3150.3630.4330.2860.2611.370.091.35
g/c-As67Se330.2010.3590.3670.4180.2910.2661.210.101.38
g/c-As70Se300.2200.4770.3760.3210.2830.2640.760.111.43
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shpotyuk, O.; Hyla, M.; Ingram, A.; Shpotyuk, Y.; Boyko, V.; Demchenko, P.; Wojnarowska-Nowak, R.; Lukáčová Bujňáková, Z.; Baláž, P. Nanostructured Molecular–Network Arsenoselenides from the Border of a Glass-Forming Region: A Disproportionality Analysis Using Complementary Characterization Probes. Molecules 2024, 29, 3948. https://doi.org/10.3390/molecules29163948

AMA Style

Shpotyuk O, Hyla M, Ingram A, Shpotyuk Y, Boyko V, Demchenko P, Wojnarowska-Nowak R, Lukáčová Bujňáková Z, Baláž P. Nanostructured Molecular–Network Arsenoselenides from the Border of a Glass-Forming Region: A Disproportionality Analysis Using Complementary Characterization Probes. Molecules. 2024; 29(16):3948. https://doi.org/10.3390/molecules29163948

Chicago/Turabian Style

Shpotyuk, Oleh, Malgorzata Hyla, Adam Ingram, Yaroslav Shpotyuk, Vitaliy Boyko, Pavlo Demchenko, Renata Wojnarowska-Nowak, Zdenka Lukáčová Bujňáková, and Peter Baláž. 2024. "Nanostructured Molecular–Network Arsenoselenides from the Border of a Glass-Forming Region: A Disproportionality Analysis Using Complementary Characterization Probes" Molecules 29, no. 16: 3948. https://doi.org/10.3390/molecules29163948

Article Metrics

Back to TopTop